You are on page 1of 231

OREGON DEPARTMENT OF TRANSPORTATION

Technical Services Branch


Environmental Section

Geo-

Geotechnical
Design Manual
Volume 1
0

Oregon Department of Transportation

Geotechnical Design Manual

Technical Services Branch


Geo-Environmental Section
Engineering and Asset Management Unit
355 Capitol St NE Rm 318
Salem OR 97301
Phone 503.986.4200 Fax 503.986.3407

Volume 1
ii

ODOT Geotechnical Design Manual


January 2011

Table of Contents
Oregon Department of Transportation......................................................................................ii
Introduction........................................................................................................................ 1-1
1.1
General...................................................................................................................... 1-1
1.1.1
Acknowledgments................................................................................................1-1
1.2
Manual Review and Comment Process......................................................................1-2
1.2.1
Manual Revision Procedure.................................................................................1-2
1.3
ODOT Geotechnical Organization..............................................................................1-3
1.4
Location of Existing Project Information.....................................................................1-4
1.5
Consultant Contracting for Geotechnical Work...........................................................1-4
2 Project Geotechnical Planning...........................................................................................2-1
2.1
General...................................................................................................................... 2-1
2.1.1
Geotechnical Project Elements............................................................................2-1
2.1.2
Geotechnical Project Tasks and Workflow...........................................................2-2
2.2
Preliminary Project Planning......................................................................................2-2
2.2.1
General................................................................................................................2-2
2.2.1.1 Project Scale and Assignment of Resources...................................................2-3
2.2.2
Office Study......................................................................................................... 2-4
2.2.3
Project Stage 1....................................................................................................2-4
2.2.4
Project Stage 2....................................................................................................2-4
2.2.4.1 Existing Information and Previous Site Investigation Data...............................2-4
2.2.4.2 Construction Records......................................................................................2-5
2.2.4.3 Site History......................................................................................................2-5
2.2.4.4 Office Research for Bridge Foundations..........................................................2-6
2.2.4.5 Site Geology....................................................................................................2-7
2.2.5
Site Reconnaissance.........................................................................................2-14
2.2.5.1 General..........................................................................................................2-14
2.2.5.2 Verification of Office Study and Site Observations.........................................2-14
2.2.5.3 Preparation for Site Exploration.....................................................................2-15
2.2.5.4 Reconnaissance Documentation...................................................................2-17
2.3
References...............................................................................................................2-17
APPENDIX 2-A Geology / Geotechnical QC MATRIX......................................................2-18
3 Field Investigation.............................................................................................................3-1
3.1
Introduction................................................................................................................ 3-1
3.1.1
Established Investigation Criteria.........................................................................3-2
3.2
General Subsurface Investigation...............................................................................3-2
3.2.1
Subsurface Investigations Phases....................................................................3-3
3.2.1.1 Phase 1........................................................................................................... 3-3
3.2.1.2 Phase 2........................................................................................................... 3-3
3.2.1.3 Other Phases..................................................................................................3-4
3.3
Exploration Plan Development...................................................................................3-4
3.3.1
Exploration Plan Considerations..........................................................................3-5
3.3.1.1 Minimum Requirements for Subsurface Investigations....................................3-5
3.3.1.2 Risk Tolerance.................................................................................................3-6
3.3.1.3 Structure Sensitivity.........................................................................................3-6
3.3.1.4 Subsurface Investigation Strategy...................................................................3-7
3.3.1.5 Schedule of Subsurface Investigations............................................................3-7
1

Volume 1
iii

ODOT Geotechnical Design Manual


June 2009

3.3.1.6 Exploration Sites..............................................................................................3-8


3.3.1.7 Right-of-Way and Permits of Entry..................................................................3-9
3.3.1.8 Utility Location/Notification.............................................................................3-10
3.3.1.9 Methods for Site Access................................................................................3-12
3.4
Exploration Management and Oversight...................................................................3-15
3.5
Subsurface Exploration Requirements.....................................................................3-16
3.5.1
General..............................................................................................................3-16
3.5.2
Exploration Spacing and Layout........................................................................3-16
3.5.2.1 Spacing and Layout Strategies......................................................................3-17
3.5.2.2 Embankment and Cut Slope Explorations.....................................................3-17
3.5.2.3 Subgrade Borings..........................................................................................3-18
3.5.2.4 Tunnel and Trenchless Pipe Installation Borings...........................................3-18
3.5.2.5 Structure-Specific Borings.............................................................................3-19
3.5.2.6 Critical-Area Investigations............................................................................3-24
3.5.2.7 Landslides.....................................................................................................3-24
3.5.3
Exploration Depths.............................................................................................3-26
3.5.3.1 Termination Depths........................................................................................3-26
3.5.3.2 Embankment and Cut Slope Exploration Depths...........................................3-26
3.5.3.3 Subgrade Borings..........................................................................................3-27
3.5.3.4 Tunnel and Trenchless Pipe Installation Borings...........................................3-27
3.5.3.5 Structure-Specific Borings.............................................................................3-27
3.5.3.6 Critical-Area Investigations............................................................................3-28
3.5.3.7 Landslides.....................................................................................................3-28
3.5.4
Sampling Requirements.....................................................................................3-29
3.5.5
Sampling Methods.............................................................................................3-29
3.5.5.1 Standard Penetration Testing........................................................................3-30
3.5.5.2 Thin-Walled Undisturbed Tube Sampling.......................................................3-30
3.5.5.3 Rock Coring...................................................................................................3-31
3.5.5.4 Bulk Sampling...............................................................................................3-32
3.5.6
Sample Disposition............................................................................................3-32
3.5.7
Exploration Survey Requirements......................................................................3-32
3.6
Subsurface Exploration Methods..............................................................................3-32
3.6.1
General..............................................................................................................3-32
3.6.2
Test Boring Methods..........................................................................................3-33
3.6.2.1 Methods Generally Not Used.........................................................................3-33
3.6.2.2 Auger Borings................................................................................................3-34
3.6.2.3 Rotary Drilling................................................................................................3-35
3.6.2.4 Rock Coring...................................................................................................3-37
3.6.2.5 Vibratory or Sonic Drilling..............................................................................3-38
3.6.2.6 Becker Hammer Drilling.................................................................................3-39
3.6.2.7 Supplemental Drilling/Exploration Applications..............................................3-39
3.6.3
Alternative Exploration Methods and Geophysical Surveys...............................3-41
3.7
Geotechnical Instrumentation...................................................................................3-42
3.7.1
General Instrumentation and Monitoring.........................................................3-42
3.7.2
Purposes of Geotechnical Instrumentation........................................................3-42
3.7.2.1 Site Investigation and Exploration..................................................................3-42
3.7.2.2 Design Verification.........................................................................................3-42
3.7.2.3 Construction and Quality Control...................................................................3-43
3.7.2.4 Safety and Legal Protection...........................................................................3-43
3.7.2.5 Performance..................................................................................................3-43
3.7.3
Criteria for Selecting Instruments.......................................................................3-43
Volume 1
iv

ODOT Geotechnical Design Manual


June 2009

3.7.3.1 Automatic Data Acquisition Systems (ADAS)................................................3-44


3.7.3.2 Instrument Use and Installation.....................................................................3-45
3.7.3.3 Inclinometers.................................................................................................3-45
3.7.3.4 Piezometers..................................................................................................3-46
3.7.3.5 Other Instruments..........................................................................................3-47
3.8
Environmental Protection during Exploration............................................................3-47
3.8.1
Protection of Fish, Wildlife, and Vegetation........................................................3-47
3.8.2
Forestry Protection.............................................................................................3-48
3.8.3
Wetland Protection.............................................................................................3-48
3.8.4
Cultural Resources Protection...........................................................................3-48
3.9
REFERENCES.........................................................................................................3-49
Appendix 3-A Permit of Entry Form.....................................................................................3-50
Appendix 3-B Utility Notification Worksheet........................................................................3-51
4 Soil and Rock Classification and Logging..........................................................................4-1
4.1
General...................................................................................................................... 4-1
5 Engineering Properties of Soil and Rock...........................................................................5-1
5.1
General...................................................................................................................... 5-1
5.2
Influence of Existing and Future Conditions on Soil and Rock Properties..................5-2
5.3
Methods of Determining Soil and Rock Properties.....................................................5-2
5.4
In-Situ Field Testing....................................................................................................5-3
5.4.1
Correction of Field SPT Values............................................................................5-3
5.5
Laboratory Testing of Soil and Rock...........................................................................5-4
5.5.1
Quality Control for Laboratory Testing..................................................................5-4
5.5.2
Developing the Testing Plan.................................................................................5-4
5.6
Engineering Properties of Soil....................................................................................5-5
5.6.1
Laboratory Performance Testing..........................................................................5-5
5.6.1.1 Disturbed Shear Strength Testing....................................................................5-5
5.6.1.2 Other Laboratory Tests....................................................................................5-6
5.6.2
Correlations to Estimate Engineering Properties of Soil.......................................5-6
5.7
Engineering Properties of Rock..................................................................................5-7
5.8
Final Selection of Design Values................................................................................5-8
5.8.1
Overview.............................................................................................................. 5-8
5.8.2
Data Reliability and Variability..............................................................................5-8
5.8.3
Final Property Selection.......................................................................................5-9
5.8.4
Development of the Subsurface Profile................................................................5-9
5.8.5
Selection of Design Properties for Engineered Materials...................................5-10
5.8.5.1 Borrow Material.............................................................................................5-10
5.8.5.2 Select Granular Backfill..................................................................................5-11
5.8.5.3 Select Stone Backfill......................................................................................5-11
5.8.5.4 Stone Embankment Material.........................................................................5-12
5.8.5.5 Wood Fiber....................................................................................................5-12
5.8.5.6 Geofoam........................................................................................................5-12
5.9
References...............................................................................................................5-13
6 Seismic Design................................................................................................................ 6-14
6.1.1
Background........................................................................................................6-15
6.1.2
Responsibility of the Geotechnical Designer......................................................6-16
6.2
Seismic Design Performance Requirements............................................................6-17
6.2.1
New Bridges......................................................................................................6-17
6.2.2
Bridge Widenings...............................................................................................6-18
6.2.3
Bridge Abutments and Retaining Walls..............................................................6-18
6.2.4
Embankments and Cut Slopes...........................................................................6-19
Volume 1
v

ODOT Geotechnical Design Manual


June 2009

6.3
Ground Motion Parameters......................................................................................6-19
6.3.1
Site Specific Probabilistic Seismic Hazard Analysis...........................................6-20
6.3.2
Magnitude and PGA for Liquefaction Analysis....................................................6-21
6.3.3
Deaggregation of Seismic Hazard......................................................................6-21
6.4
Site Characterization for Seismic Design..................................................................6-23
6.4.1
Subsurface Investigation for Seismic Design.....................................................6-23
6.5
Geotechnical Seismic Design Procedures................................................................6-31
6.5.1
Design Ground Motion Data...............................................................................6-34
6.5.1.1 Development of Design Ground Motion Data................................................6-34
6.5.1.2 AASHTO General Procedure.........................................................................6-34
6.5.1.3 Response Spectra and Analysis for Liquefied Soil Sites................................6-34
6.5.1.4 Ground Response Analysis...........................................................................6-35
6.5.1.5 Selection of Time Histories for Ground Response Analysis...........................6-36
6.5.1.6 Ground Motion Parameters for Other Structures...........................................6-38
6.5.1.7 Bedrock versus Ground Surface Acceleration...............................................6-39
6.5.2
Liquefaction Analysis.........................................................................................6-39
6.5.2.1 Liquefaction Design Policies..........................................................................6-40
6.5.2.2 Methods to Evaluate Liquefaction Potential...................................................6-41
6.5.2.3 Liquefaction Induced Settlement...................................................................6-44
6.5.2.4 Residual Strength Parameters.......................................................................6-46
6.5.3
Slope Stability and Deformation Analysis...........................................................6-47
6.5.3.1 Pseudo-static Analysis...................................................................................6-47
6.5.3.2 Deformation Analysis.....................................................................................6-48
6.5.4
Settlement of Dry Sand......................................................................................6-51
6.5.5
Liquefaction Effects on Structure Foundations...................................................6-52
6.5.5.1 Bridge Approach Fills.....................................................................................6-52
6.5.5.2 General Liquefaction Policies Regarding Bridge Foundations.......................6-52
6.5.5.3 Lateral Spread / Slope Failure Loads on Structures......................................6-54
6.5.5.4 Displacement Based Approach......................................................................6-54
6.5.5.5 Force Based Approaches..............................................................................6-55
6.5.6
Mitigation Alternatives for Lateral Spread...........................................................6-56
6.6
Input for Structural Design........................................................................................6-58
6.6.1
Foundation Springs............................................................................................6-58
6.6.1.1 Shallow Foundations.....................................................................................6-58
6.6.1.2 Deep Foundations.........................................................................................6-59
6.6.1.3 Downdrag Loads on Structures.....................................................................6-59
6.7
References...............................................................................................................6-60
Appendix 6-A...................................................................................................................... 6-65
Appendix 6-B: Example of Smoothed Response Spectra...................................................6-67
Appendix 6-C: ODOT Liquefaction Mitigation Procedures...................................................6-68
7 Slope Stability Analysis......................................................................................................7-1
7.1
General...................................................................................................................... 7-1
7.2
Development of Design Parameters and Input Data for Slope Stability Analysis........7-1
7.3
Design Requirements.................................................................................................7-2
7.4
Resistance Factors and Safety Factors for Slope Stability Analysis...........................7-3
7.5
References.................................................................................................................7-3
8 Foundation Design............................................................................................................8-1
8.1
General...................................................................................................................... 8-1
8.2
Project Data and Foundation Design Requirements...................................................8-1
8.3
Field Exploration for Foundations...............................................................................8-4
8.4
Field and Laboratory Testing for Foundations.............................................................8-4
Volume 1
vi

ODOT Geotechnical Design Manual


June 2009

8.5
Material Properties for Design....................................................................................8-5
8.6
Bridge Approach Embankments.................................................................................8-5
8.6.1
Abutment Transitions...........................................................................................8-6
8.6.2
Overall Stability....................................................................................................8-6
8.7
Foundation Selection Criteria.....................................................................................8-7
8.8
Overview of LRFD for Foundations............................................................................8-9
8.9
Foundation Design Policies......................................................................................8-10
8.9.1
Downdrag Loads................................................................................................8-10
8.9.2
Scour Design.....................................................................................................8-10
8.9.3
Seismic Design..................................................................................................8-11
8.10 Soil Loads on Buried Structures...............................................................................8-12
8.11 Spread Footing Design.............................................................................................8-12
8.11.1 Nearby Structures..............................................................................................8-12
8.11.2 Service Limit State Design of Footings..............................................................8-13
8.12 Driven Pile Foundation Design.................................................................................8-13
8.12.1 Required Pile Tip Elevation................................................................................8-14
8.12.2 Pile Drivability Analysis and Wave Equation Usage...........................................8-15
8.12.3 Pile Setup and Restrike.....................................................................................8-15
8.12.4 Driven Pile Types, and Sizes.............................................................................8-15
8.12.5 Extreme Event Limit State Design......................................................................8-17
8.12.5.1 Scour Effects on Pile Design ......................................................................8-17
8.13 Drilled Shaft Foundation Design...............................................................................8-21
8.13.1 Nearby Structures..............................................................................................8-22
8.13.2 Scour.................................................................................................................8-22
8.13.3 Extreme Event Limit State Design of Drilled Shafts...........................................8-22
8.14 Micropiles................................................................................................................. 8-23
8.15 References...............................................................................................................8-23
9 Embankments Analysis and Design................................................................................9-1
9.1
General...................................................................................................................... 9-1
9.2
Design Considerations...............................................................................................9-2
9.2.1
Typical Embankment Materials and Compaction.................................................9-2
9.2.1.1 All-Weather Embankment Materials................................................................9-2
9.2.1.2 Durable and Non-Durable Rock Materials.......................................................9-2
9.2.1.3 Earth Embankments.........................................................................................9-3
9.2.2
Embankment Stability Assessment......................................................................9-3
9.2.2.1 Safety Factors.................................................................................................9-3
9.2.2.2 Strength Parameters.......................................................................................9-4
9.2.3
Embankment Settlement Assessment..................................................................9-4
9.2.3.1 Settlement Analysis.........................................................................................9-4
9.2.3.2 Analytical Tools................................................................................................9-5
9.3
Stability Mitigation......................................................................................................9-5
9.3.1
Staged Construction............................................................................................9-5
9.3.2
Base Reinforcement............................................................................................9-5
9.3.3
Ground Improvement...........................................................................................9-6
9.3.4
Lightweight Fills...................................................................................................9-6
9.3.5
Toe Berms and Shear keys..................................................................................9-6
9.4
Settlement Mitigation..................................................................................................9-6
9.4.1
Acceleration Using Wick Drains...........................................................................9-6
9.4.2
Acceleration Using Surcharges............................................................................9-6
9.4.3
Lightweight Fills...................................................................................................9-7
9.4.4
Subexcavation.....................................................................................................9-7
Volume 1
vii

ODOT Geotechnical Design Manual


June 2009

9.5
References.................................................................................................................9-8
10
Soil Cuts - Analysis and Design...................................................................................10-1
10.1 General.................................................................................................................... 10-1
10.1.1 Design Parameters............................................................................................10-2
10.2 Soil Cut Design.........................................................................................................10-2
10.2.1 Design Approach and Methodology...................................................................10-2
10.2.2 Seepage Analysis and Impact on Design...........................................................10-3
10.2.3 Surface and Subsurface Drainage Considerations and Design..........................10-3
10.2.4 Stability Improvement Techniques......................................................................10-4
10.2.5 Erosion and Piping Considerations....................................................................10-5
10.2.6 Sliver Cuts.........................................................................................................10-5
10.3 References...............................................................................................................10-1
11
Ground Improvement...................................................................................................11-1
11.1 General..................................................................................................................... 11-1
11.2 Development of Design Parameters and Other Input Data for Ground Improvement
Analysis............................................................................................................................... 11-2
11.3 Design Requirements...............................................................................................11-2
11.4 References...............................................................................................................11-4
12
Rock Cuts Analysis, Design and Mitigation...............................................................12-1
12.1 General.................................................................................................................... 12-1
12.2 ODOT Rock Slope Design Policy.............................................................................12-1
12.2.1 Rock Slope Design............................................................................................12-1
12.2.2 Rockslope Fallout Areas....................................................................................12-1
12.2.3 Benches............................................................................................................. 12-2
12.2.4 Rock Slope Stabilization and Rockfall Mitigation Techniques.............................12-3
12.3 Rockslope Stability Analysis.....................................................................................12-3
12.4 Design Guidelines....................................................................................................12-3
12.4.1 Geologic Investigation and Mapping..................................................................12-3
12.4.2 Analysis and Design..........................................................................................12-3
12.4.3 Construction Issues...........................................................................................12-4
12.4.4 Blasting Consultant............................................................................................12-4
12.4.5 Gabion Wire Mesh Slope Protection/ Cable Net Slope Protection.....................12-4
12.4.6 Rock Reinforcing Bolts and Rock Reinforcing Dowels.......................................12-4
12.4.7 Proprietary Rockfall Net Systems.......................................................................12-5
12.5 Standard Details.......................................................................................................12-5
12.6 Specifications...........................................................................................................12-5
12.6.1 Blasting..............................................................................................................12-5
12.6.2 Rockslope Mitigation Methods...........................................................................12-6
12.7 References...............................................................................................................12-1
13
Slope Stability Analysis................................................................................................13-1
13.1 General.................................................................................................................... 13-1
14
Geosynthetic Design....................................................................................................14-1
14.1 General.................................................................................................................... 14-1

Volume 1
viii

ODOT Geotechnical Design Manual


June 2009

List of Figures
Figure 6-1. Shear modulus reduction and damping ratio curves for sand (EPRI, 1993)..........6-28
Figure 6-2. Variation of G/Gmax vs. cyclic shear strain for fine grained soils (redrafted from
Vucetic and Dobry, 1991)............................................................................................6-29
Figure 6-3. Equivalent viscous damping ratio vs. cyclic shear strain for fine grained soils
(redrafted from Vucetic and Dobry, 1991)....................................................................6-29
Figure 6-4. Residual undrained shear strength for liquefied soils as a function of SPT blow
counts (Seed and Harder, 1990 and Idriss, 2003).......................................................6-30
Figure 6-5. Estimation of residual strength ratio from SPT resistance (Olson and Stark, 2002).. 630
Figure 6-6. Variation of residual strength ratio with SPT resistance and initial vertical effective
stress using Kramer-Wang model (Kramer, 2008).......................................................6-31
Figure 6-7. General Geotechnical Seismic Design Procedures..............................................6-33
Figure 6-8. Magnitude Scaling Factors Derived by Various Investigators (redrafted from 1996
NCEER Workshop Summary Report).........................................................................6-44
Figure 6-9. Liquefaction induced settlement estimated using the Tokimatsu & Seed procedure
(redrafted from Tokimatsu and Seed, 1987)................................................................6-45
Figure 6-10. Liquefaction induced settlement estimated using the Ishihara and Yoshimine
procedure. (redrafted from Ishihara and Yoshimine, 1992)..........................................6-46
Figure 6-11. The Makdisi-Seed procedure for estimating the range of permanent seismically
induced slope deformation as a function of the ratio of yield acceleration over maximum
acceleration (redrafted from Makdisi and Seed, 1978)................................................6-50
Figure 6-12. Lateral Extent of Ground Improvement for Liquefaction Mitigation.....................6-57
Figure 8-1. Example where footing contributes to instability of slope (left figure) vs. example
where footing contributes to stability of slope (right figure)............................................8-7
Figure 8-2. Design of pile foundations for scour......................................................................8-18
Figure 8-3. Design of pile foundations for liquefaction downdrag (WSDOT, 2006)................8-20

List of Tables
Table 2-1. Geology / Geotechnical Matrix Checklist QC Check #1 Scoping.........................2-18
Table 2-2. Geology / Geotechnical Matrix Checklist QC Check #2 Scope of Work..............2-19
Table 2-3. Geology / Geotechnical Matrix Checklist QC Check # 3 EIS...............................2-20
Table 2-4. Geology / Geotechnical Matrix Checklist QC Check # 4 Concept.......................2-21
Table 2-5. Geology / Geotechnical Matrix Checklist QC Check #5 Exploration Plan (10%
TS&L).......................................................................................................................... 2-22
Table 2-6. Geology / Geotechnical Matrix Checklist QC Check #6 2/3 TS&L)......................2-23
Table 2-7. Geology / Geotechnical Matrix Checklist QC Check #7 Preliminary Plans..........2-29
Table 2-8. Geology / Geotechnical Matrix Checklist QC Check #8 Advanced Plans............2-32
Table 2-9. Geology / Geotechnical Matrix Checklist QC Check #9 Final Plans....................2-33
Table 3-1. Tunneling and Trenchless Pipe Installation Recommendations..............................3-19
Table 3-2. Specific field investigation requirements................................................................3-23
Table 5-1. Correlation of SPT N values to drained friction angle of granular soils (modified after
Bowles, 1977)............................................................................................................... 5-6
Table 6-1. Summary of site characterization needs and testing considerations for seismic design
(adapted from Sabatini, et al., 2002)...........................................................................6-25
Table 6-2. Correlations for estimating initial shear modulus (Kavazajjian, et al., 1997)...........6-27
Table 8-1. Summary of information needs and testing considerations (modified after Sabatini, et.
al. 2002)......................................................................................................................... 8-2

Volume 1
ix

ODOT Geotechnical Design Manual


January 2011

Volume 1
x

ODOT Geotechnical Design Manual


January 2011

Chapter

Introduction

General

The ODOT Geotechnical Design Manual (GDM) establishes standard policies and procedures
regarding geotechnical work performed for ODOT. The manual covers geotechnical investigations,
analysis, design and reporting for earthwork and structures for highways. The purpose of the
geotechnical investigation and design recommendations is to furnish information for an optimum
design which will minimize over-conservatism as well as to minimize under-design and the resulting
failures commonly and mistakenly attributed to unforeseen conditions.
It is to be understood that any geotechnical investigation and design will leave certain areas
unexplored. Further, it must also be understood that it would be impractical to provide a rigid set of
specifications for all possible cases. Therefore, this manual will not address all subsurface problems
and leaves many areas where individual engineering judgment must be used. It is intended that the
procedures discussed in this manual will establish a reasonable and uniform set of policies and
procedures while maintaining sufficient flexibility to permit the application of engineering analysis to
the solution of geotechnical problems.
This manual references publications which present specific engineering design and construction
procedures or laboratory testing procedures. Each chapter contains a listing of associated
references for the subject area of the chapter. Among the commonly referenced materials are the
publications of the American Association of State Highway Transportation Officials (AASHTO), the
Federal Highway Administration (FHWA), and the American Society for Testing and Materials
(ASTM). Relative to testing and design procedures, the methods presented by AASHTO and FHWA
are often followed.
Figures presented in the manual have been redrafted from the original published figures. The figures
in this GDM are only to be used for illustrative purposes and should not be used for design.

1.1.1

Acknowledgments

This ODOT Geotechnical Design Manual is a completely new manual and is the product of the
combined efforts of the personnel in the HQ Engineering and Asset Management Unit. Thanks for
their work and appreciation for their contributions are extended here. Continued work is required to
edit and update the manual and their help will be appreciated in the future. An additional thanks and
Volume 1
1

ODOT Geotechnical Design Manual


January 2011

acknowledgement is given to Tony Allen of WSDOT for permission to use, en masse, whole sections,
paragraphs, and even an entire chapter or two in the development of this manual.
The completion of the WSDOT Geotechnical Design Manual in September 2005 provided the spark
and impetus for ODOT to finally, after many years of wishing it, produce a Geotechnical Design
Manual worthy of the importance of geotechnical design on highway projects.

Manual Review and Comment Process

The ODOT Engineering and Asset Management Unit of the Geo-Environmental Section is
responsible for the publication and modification of this manual. Any comments or questions about
the ODOT Geotechnical Design Manual should be directed to:

Paul Wirfs, P.E., Unit Manager


Engineering and Asset Management Unit
Geo-Environmental Section
Oregon Department of Transportation
355 Capitol St NE, Rm 301
Salem, OR 97301
503-986-4200

1.2.1

Manual Revision Procedure

It is intended that the GDM will be continually updated as required to clarify geotechnical practice in
ODOT and include new information. Revisions and submittals from all users of the GDM, both inside
ODOT and Consultants, are encouraged. The following revision procedure should be used:
1. Define the problem
Discuss the suggestion or revision of the GDM with others that have a stake in the outcome.
If it is agreed that the item should be proposed, develop a written proposal. Changes to
design policy, design practice or procedure can have wide ranging effects and will affect
some or all of those involved in the preparation of contract documents for ODOT.
2. Put it in writing
Research and develop a written proposal using the three general subject headings:

Problem Statement

Analysis/ Research Data

Proposal

Check the finished product by reviewing the following guiding comments:

The existing problem is clearly stated

Research and analysis of the problem and potential solution are thorough and
understandable

The proposed solution is well thought out, is supported by facts and solves the
problem. Has the impact on other areas been considered? Have the details been
coordinated with other units or organizations that may be affected?

Volume 1
2

ODOT Geotechnical Design Manual


January 2011

No questions remain that need to be answered before implementation

3. Review and Approval


After reviewing the written proposal for completeness, the Engineering and Asset
Management Unit will either:

Accept, without further review, manual corrections for inclusion in the GDM, or

Distribute a copy with the due date and a Geotechnical Design Practice Approval
Form for review and comments.

After reviewing the returned Geotechnical Design Practice Approval Forms, the Engineering
and Asset Management Unit will do one of the following:

Proposals approved for revision of the GDM will be implemented in a technical


bulletin and will be placed into the next upcoming version of the GDM, or

Proposals needing more research or clarification will be returned to the originator for
revision and resubmittal.

Regardless of whether or not a proposal is accepted, the Engineering and Asset


Management Unit will reply in writing to the person making the submittal.
4. Implementation of Approved Revision
After a proposal has final approval, a revised GDM page will be prepared for inclusion into the
manual. A vertical line in the right-hand margin of a revised page indicated that the text has
been revised or added. The word REVISION and the year are printed in the bottom margin
to the right. This system is similar to that used by AASHTO to revise its Standard
Specifications.
Proposals will be incorporated electronically into the GDM on the ODOT Geo-Environmental
web page as soon as practical.

ODOT Geotechnical Organization

The functions of geotechnical design in ODOT are generally managed and performed within the 5
region offices. Tech Centers within each region are staffed with Geotechnical Engineers, Engineering
Geologists, Hydraulics Engineers, and HazMat specialists. The geotechnical design, construction
and maintenance support may be performed in-house or contracted out to specialty consultants. The
ODOT Headquarters Engineering and Asset Management Unit provides on-call geotechnical design
assistance and review, training and software, coordination of section initiatives, and other functions
involving development of standards and practices for geotechnical work. Material source and
aggregate material program needs are also a function of the headquarters unit. Currently, a
significant portion of geotechnical work is being done under the OTIA Bridge Program and is being
managed by the Oregon Bridge Delivery Partners (OBDP). OBDP manages and reviews consultant
geotechnical work for design-bid-build and design-build OTIA III highway projects.

Volume 1
3

ODOT Geotechnical Design Manual


January 2011

For reference purposes, the following are the contacts and location of the managers for each of the
groups tasked with geotechnical work in ODOT:
Region

Manager

Phone

Fax

1 Portland

Tom Braibish

503-731-8290

503-731-3164

2 Salem

Bernie Kleutsch

503-986-2646

503-986-2622

3 Roseburg

Pete Castro

541-957-3603

541-957-3604

4 Bend

Randy Davis

541-388-6334

541-385-0476

5 LaGrande

Mark Hanson

541-963-1361

541-963-9079

HQ Geotech Salem

Paul Wirfs

503-986-3526

541-986-3407

HQ Bridge Salem

Bruce Johnson

503-986-3344

541-986-3407

Location of Existing Project Information

In general, the regional offices keep file information on past projects. The first inquiry into project
geotechnical information should be to the appropriate region Geotechnical office. In addition, project
information for past projects involving geotechnical analysis and design has been archived and
stored in the ODOT Salem Airport Road complex. A database listing of the projects archived is
located in the HQ Salem Engineering and Asset Management office in Salem. The Salem Bridge
Engineering Section keeps pile record books for past projects where pile driving was performed. In
addition, bridge archives are available that include Foundation Reports, boring logs, as-constructed
bridge plans and foundation data sheets. Inquiries regarding bridge foundation records and archives
should be directed to the HQ Bridge Section office.

Consultant
Contracting
Geotechnical Work

for

ODOT has a set of specialty consultants retained to perform geotechnical work as needed. The
current list of geotechnical consultants can be obtained from the ODOT Procurement Office (OPO).
A Scope of Work Template has been developed for use by those needing to have a consultant
perform geotechnical work and is located on the ODOT Geo-Environmental website.

Volume 1
4

ODOT Geotechnical Design Manual


January 2011

Chapter

Project Geotechnical Planning

General

General geotechnical planning for projects with significant grading, earthworks, and structure
foundations, from the earliest project concept plan through final project design are addressed in this
chapter. Detailed geotechnical exploration and testing requirements for individual design are covered
in detail in Chapter 3, Chapter 4 and Chapter 5. This chapter also provides direction for
geotechnical project definition and creation of the subsurface exploration plan for the project design
phases. General guidelines for subsurface investigations are provided in Chapter 3 in addition to
specific guidelines regarding the number and types of explorations for project design of specific
geotechnical features.
The success of a project is directly related to the early involvement of the geotechnical designers in
the design process. For larger projects that involve an Environmental Impact Statement (EIS), the
geotechnical designer needs to be involved with the assessment of various options or corridor
selections. Ideally, for all projects, the geotechnical designer will be involved during the first scoping
efforts. At this point, a study of the project concept is begun by gathering all existing site data and
determining the critical features of the project. This information can then be presented at the project
kick-off meeting and/or scoping trip. The project scoping trip is a valuable opportunity to introduce
the roadway and structural designers, and project leaders to the geologic/geotechnical issues that
are expected to impact the project. Continued good communication between the geotechnical
designer and the project leader and project team is vital.

2.1.1

Geotechnical Project Elements

All proposed project scopes should be reviewed by an engineering geologist and/or geotechnical
engineer for a determination of the project elements (if any) that require a geologic investigation and
geotechnical design. This allows the geotechnical designers to begin formulating a prospective
scope of work and budget estimate. There are common project elements that are always the subject
of a geotechnical investigation and design such as bridge foundations and landslide mitigations, and
there are project elements that, depending on the site history and underlying geology, may or may
not need investigation and design, or may require different levels of effort. The geotechnical
designers will be able to determine the level of effort based on their own or others knowledge and
experience of the site to make these judgments. Because of the underlying site conditions, elements
that generally dont warrant geotechnical design for most sites may require it at others. Conversely,
Volume 1
1

ODOT Geotechnical Design Manual


April 2010

investigation and design efforts may be scaled back or eliminated at other sites due to known
favorable conditions, and the significance of the project feature. It is the responsibility of the
geotechnical designers to make these decisions.
The common project elements on transportation projects that are the subject of engineering geologic
investigation and geotechnical design for construction are:

Structure Foundations (bridges, viaducts, pumping stations, sound walls, buildings, etc.)

Retaining walls over 4 in height as measured from the base of the wall footing to the top of
the wall and any wall with a foreslope or backslope

Cuts and embankments over 4 in height

Tunnels and underground structures

Poles, masts and towers

Culverts, pipes and conduits

This last group of elements, culverts, pipes and conduits, exemplify the broad range of design and
investigation that may occur on any project. A 24 culvert replacement at a depth of 3 feet below a
proposed roadway alignment would normally require the hand-collection of soil samples from the
pipe location, submittal of those samples to the laboratory for chemical properties testing, and
forwarding the results to the project designer for selection of the appropriate pipe materials for that
location. If however, that same culvert was to be installed under a large, existing embankment while
under traffic using trenchless methods, then the required investigation and design effort would be
close to what is required for a tunnel or underground structure.

2.1.2

Geotechnical Project Tasks and Workflow

The expected milestones for geotechnical input on projects and the review of geotechnical work is
outlined in APPENDIX 2-A Geology / Geotechnical QC MATRIX, and the Project Flowchart.
Certain project checkpoints and tasks may be added or eliminated based on the project scope and/or
requirements. Each individual project prospectus should be consulted to determine which tasks and
QC checkpoints will apply.

2.2.1

Preliminary Project Planning


General

The creation of an efficient geologic/geotechnical investigation and identification of fatal flaws or


critical issues that could impact design and construction as early in project development as possible
is essential. Use the maximum amount of effort to obtain the greatest amount of information as early
in each phase of investigation as possible so that each successive phase can capitalize on the
information previously gathered. The result is a more thorough and cost-effective geologic and
geotechnical investigation program.
Projects with a small number of defined structure locations or limited earthwork typically do not
require numerous phases of investigation. Such projects normally proceed through an initial
background study, site reconnaissance and ensuing subsurface exploration at the TS&L phase.
Larger projects in contrast, will usually benefit from a phased sequence of field exploration. The
geologic/geotechnical investigation will occur as a reconnaissance-level examination and preliminary
subsurface exploration during the Field Survey phase of the project. More detailed, site-specific
Volume 1
Manual

ODOT Geotechnical Design


2

April 2010

exploration is accomplished later as the project develops through the TS&L and Approved Design
phases.
Phased subsurface exploration is beneficial because:

Phased subsurface exploration allows information to be obtained in the early stage of the
project that can be used to focus the exploration plan for the more detailed design stages.
This is where previously gained information can be used to maximize the efficiency of the
final exploration, and to assure that previously identified geotechnical problems and/or
geologic hazards are thoroughly investigated and characterized.

Additionally, the Exploration Plan can be more clearly defined and easier to manage. In this
regard, the number of borings, their depths, and laboratory testing programs can be
determined in advance of actual mobilization of equipment to a project area.

For most projects, mobilization costs for exploration equipment are high, so efforts should be made to
reduce the number of subsurface investigation phases whenever practical. However, the site
location, project objectives, and other factors will necessarily influence the investigation phases and
mobilizations. Some of the additional factors to consider are site access, availability of specialized
equipment, environmental restrictions, safety issues, and traffic control.
To economize field investigations and provide contingencies for ongoing project changes, consider
the following:

A substantial amount of background study should take place prior to mobilization to a project
site. The information derived from this research provides a basis for the design of the
Exploration Plan and help focus the on-site investigation.

In addition, all resources used in the development of the background study should be
organized and documented in such a manner that another geotechnical designer would be
able to continue the project without going back to the beginning to get the same information.
Keep a list of all documents used in the background study, such as field notes and sketches
from initial site reconnaissance, reports or investigations from previous or nearby site
investigations, and other published literature.

Any critical issues such as geologic hazards, problem materials or conditions, or


contamination identified during the initial study should be clearly documented and highlighted
throughout the project to avoid any surprises later on in the design or construction phases.

2 . 2 . 1 . 1 P r o j e c t S c a l e an d As s i g n m e n t o f R e so u r c e s
Geotechnical designers should use their professional judgment with respect to the scope, scale, and
amount of resources to utilize during preliminary project studies. Larger projects obviously
necessitate a greater effort in the early examination of background materials such as previous
reports for an area, maps, published literature, aerial photographs and other remote sensing.
Even the smallest bridge replacement or grading project, background study is just as important, and
although of a smaller scale, should be carried out with the same diligence as a similar study for a
major realignment. A thorough and expedient background study is essential for these smaller
projects since unforeseen conditions and additional unplanned field investigations are much more
difficult to absorb in a smaller project budget. It follows that for a larger project; a more thorough
background investigation is warranted since unforeseen conditions can have a compounding effect
during design and construction that may impact even the most generously funded projects.
The amount of background research needed for a project is usually unknown until the study begins
and the potential site conditions are assessed to some degree. It is up to the geotechnical designer
Volume 1
Manual

ODOT Geotechnical Design


3

April 2010

to determine the amount of background study needed and the cost-benefit of such studies with
respect to the project design.
Using Remote Sensing and Existing Information
Ordering new remote sensing studies to assess surrounding landforms is probably not necessary for
in-kind bridge replacement projects unless some special conditions are observed during the field or
office study. However, failure to procure and study a set of aerial photographs along a proposed
realignment would probably be somewhat negligent. Project background studies for major
realignment projects and landslide mitigations typically make more use of remote sensing and
published literature while replacement and modernization projects will rely more heavily on previous
site studies and reports.

2.2.2

Office Study

The foremost objectives of initial office study are 1) early identification of critical issues that will affect
the projects scope, schedule, or budget, and 2) efficiently plan detailed site studies and formulate a
subsurface investigation program.

2.2.3

Project Stage 1

The first stage of any project should begin with a review of the published and available unpublished
literature to gain a thorough understanding of the existing site conditions and composition. Such an
understanding includes knowledge of the geologic processes that have been the genesis of, or have
in some way affected the project site. The site geomorphology should receive the most scrutiny from
the geotechnical designer since characteristic landforms are created by specific geologic processes,
and composed of particular materials. The site geomorphology, coupled with the literature and
results of previous studies, will aid the geotechnical designer in predicting what materials will be
encountered, and how they will be distributed across the site.

2.2.4

Project Stage 2

The second stage of a project involves the detailed examination of the proposed project components
and in particular, the geotechnical elements. This includes an appraisal of the project prospectus as
well as any conceptual or preliminary plans available from the roadway designer or project leader.
The project geotechnical features such as bridge foundations; earth retaining structures, cuts,
embankments and any other earthworks should be identified and located. Once the project
geotechnical features are recognized, they can then be analyzed with respect to the background
information previously collected.

2 . 2 . 4 . 1 E x i s t i n g I n f o r m a t i o n a n d P r e vi o u s S i t e
I n ve s t i g a t i o n D a t a
Current transportation projects take place almost exclusively on or near existing routes, for which a
considerable amount of subsurface information already exists, in most cases. Since many
transportation projects take place in urban areas, additional information may also be available from
other nearby public works projects and private developments involving structures and earthworks.
Local agencies may possess subsurface information for their projects as well as data provided by
consultants.
Subsurface information collected for ODOT projects resides in the region geology office in which the
data was collected. Subsurface information is collected for bridge foundations, retaining walls, cut
slopes, embankments, and landslides. Additional subsurface data has also been collected for

Volume 1
Manual

ODOT Geotechnical Design


4

April 2010

incidental structures such as sound walls, sign bridges, poles, masts and towers, and facilities such
as water tanks and maintenance buildings.
The Oregon Water Resources Department maintains a database of boring logs on its website. By
law, reports must be filed with this agency for all geotechnical holes and water, thermal, and
monitoring wells. Thus, the database is fully populated, and may be queried in many ways
geographically or by owner, number, constructor, or purpose. These logs are beneficial in rural or
remote areas with a dearth of subsurface information.
Note:
A wealth of information can be contained on the logs especially regarding groundwater and depth to
bedrock information. There is an entry for soil and rock descriptions on the reporting forms. However
this information should be used with caution since there are no standard reporting formats and thus,
the soil and rock descriptions on the Water Resources forms vary in content and accuracy.
The Oregon Department of Water Resources Database (ORWD) can be accessed at the following
location:
http://apps2.wrd.state.or.us/apps/gw/well_log/Default.aspx
In addition to the information provided on the OWRD forms, it is important to simply note the
presence of wells in the area that may be affected by the project construction. Projects involving
large cut slopes or dewatering efforts can affect the yield of nearby wells. Where this occurs, ODOT
typically includes replacement or deepening of the well as part of the Right-of-Way acquisition.

2 . 2 . 4 . 2 C o n s t r u c t i o n R e co r d s
Since most current ODOT projects are modernization, replacement, or rehabilitations of existing
transportation facilities, construction records are commonly available from various sources
throughout the agency. Such records may be in the form of as-built plans, construction reports, piledriving records, and other technical memoranda addressing specific issues and recommendations
during project construction. Locate information using:

As-built plans: As-built plans are normally located in the region office where the project was
constructed. The Geometronics Unit maintains the engineering documents in Room 29 of
the Transportation Building in Salem where mylars of project plans reside in addition to some
of the as-built plans.

Pile records: Pile record books are maintained by the headquarters office of the Bridge
Section.

Region project engineers and construction project managers that have completed previous projects
in the area should be consulted with respect to the geologic/geotechnical conditions as well as the
construction issues related to those conditions. In addition, section maintenance personnel with a
long history in an area will posses a wealth of information regarding the performance of existing
facilities, problems encountered, and repair activities that have taken place at a particular site.

2.2.4.3 Site History


Past use of a site can greatly affect the design and construction of a project and can also make a
significant impact to its timeline and budget. Typically, much of a sites background and past use will
be researched and described for a Phase I or II Environmental Site Assessment produced by the
environmental specialists or their consultants in the region geology offices. Information concerning
the development of Environmental Site Assessments and other site use resources can be found in
the HazMat Manual. Environmental Impact Statements (EIS) for previous projects in the area are
also an important and concise source of previous and current site use information. Some of the
Volume 1
Manual

ODOT Geotechnical Design


5

April 2010

remote sensing methods previously discussed may also help determine previous site use in the
absence of historic records.

Hazardous Materials
The presence of hazardous materials in the subsurface not only affects the geotechnical design, and
the construction approach to a project, but it also greatly affects how the subsurface investigation
program is carried out. For this reason itself, it becomes important for the geotechnical designer to
determine if previous use of the site, or surrounding locations could have potentially resulted in
subsurface contamination. Such uses include any facility or enterprise engaged in the production,
distribution, storage, or use of hazardous substances. Hazardous substances are defined by the
Environmental Protection Agency (EPA) in 40CFR261.31 through 261.33. In addition, the EPA
further includes as hazardous wastes, such substances with characteristics of Ignitability, Corrosively,
Reactivity, and Toxicity according to 40CFR261.21 through 261.24. For transportation projects, the
most commonly contaminated sites are those that are presently, or have previously been occupied
by service stations. However, larger manufacturing and processing sites with substantial amounts of
contamination are encountered. When geotechnical investigation must be conducted under such
conditions, significant preplanning is required not only to protect the field crew, but also to comply
with the numerous environmental regulations that govern everything from required PPE to disposal of
contaminated drill cuttings.

Previous Site Use


In addition to contaminated materials, previous site uses have the possibility of leaving behind
materials and/or conditions that can be detrimental to the construction or performance of a facility if
not properly mitigated. In this regard, deleterious fill materials such as wood waste and ash are
commonly associated with timber processing and other operations throughout the state while
reclaimed quarries may be filled with deep, unconsolidated debris and spoils. Underground mines
and tunnels are present in various locations throughout Oregon. Although uncommon, some
instances of such features unexpectedly encountered during construction have occurred. In addition
to their obvious geotechnical impacts, such features may be historic locations and thus, be protected
by Federal law.

Previous Site Occupation


In addition to previous site use, the geotechnical designer must also consider previous site
occupation. A site previously occupied by Native Americans can contain artifacts, or be of
significance to contemporaries. Such occupation may require archaeological investigation or
preservation activities by qualified personnel. It is also possible that the exploration plan, or even
significant project design changes prior to on-site geotechnical investigation will be required. Historic
sites, structures, and even trees will also be protected in some instances that will necessitate
adjustments to the proposed investigation. Clearly, much of the archaeological and historical issues
in connection with a site are outside the purview of engineering geology and geotechnical
engineering. However, the geotechnical designer must be aware of these issues to assure that field
investigation activities are compliant with the laws and regulations that protect these resources.

2.2.4.4 Office Research for Bridge Foundations


In addition to the sources of information listed above, office research for bridge foundation work
generally consists of a review of foundations for the existing structure and any other pertinent
foundation information on other nearby structures. The structure owner may have subsurface
information such as soil boring logs or as-constructed foundation information such as spread footing
elevations, pile tip elevations, or pile driving records. The HQ Bridge Section archives contain
Foundation Reports and boring logs for many bridges constructed between the mid 1960s to about
2001. Subsurface information on some earlier ODOT bridges may also be available in the Bridge
Volume 1
Manual

ODOT Geotechnical Design


6

April 2010

Section construction records. Between about 1999 and 2004, bridge foundation files, reports and
records for most bridges were stored in the Salem Geo/Hydro Section archives (now the GeoEnvironmental Section archives). Copies of these reports should also exist in the region offices.
Maintenance and construction records for existing bridge(s) should also be reviewed for information
relevant to the design and construction of the proposed structure. These records are available in the
HQ Bridge Engineering Library or from the Bridge Section Archives. As-Constructed bridge drawings
are available online, internally to ODOT through the ODOT Bridge Data System (BDS). Pile driving
record books are also available from the HQ Bridge Section.
Office research work for structure foundations typically includes (but is not limited to) gathering the
following information for the existing structure(s):

Location and structure dimensions, number of spans, year constructed

Superstructure type (e.g. RCDG, composite, steel beam)

Subsurface data (e.g. foundation reports, boring logs, data sheets, groundwater conditions,
etc.)

Type of Foundation (e.g. spread footings, piles, shafts)

Applicable as-constructed foundation information such as:

Spread footing elevation, dimensions, and design or applied load

Pile type and size, pile tip elevations or lengths, design or actual driven pile capacity and the
method used to determine capacity (resistance) (dynamic formula (ENR, Gates), wave
equation, PDA/CAPWAP)

Drilled shaft diameter, tip elevations

Construction problems (e.g., groundwater problems, boulders or other obstructions, caving,


difficult shoring/cofferdam construction).

Foundationrelated maintenance problems (e.g., approach fill or bridge settlement, scour


problems, rip rap placement, corrosion, slope stability or drainage problems)

A review of old roadway design plans, air photos, and soil and geology maps and well logs may also
be useful. Particular attention should be given to locating any existing or abandoned foundations or
underground utilities in the proposed structure location. Any obstructions or other existing conditions
that may influence the bridge design, bent layout or construction should be communicated directly to
the structural designer as soon as possible so these conditions can be taken into account in the
design of the structure.
This information should be summarized and provided in the Geotechnical Report. All applicable asconstructed drawings or boring logs for the existing structure should be included in the Geotechnical
Report Appendices.

2.2.4.5 Site Geology


The underlying geology of a project site provides important information concerning the conditions that
may be encountered during the investigation and construction phases of a project. Of equal
importance is the indication of conditions that either may not be encountered, or will require specific
procedures to determine if they do exist. Some particularly deep bedrock horizons, groundwater
surfaces, and boulders or other obstructions are examples. Certain conditions can be expected due
to the nature of the project site geology.

Volume 1
Manual

ODOT Geotechnical Design


7

April 2010

Oregon has specific geologic terrains, formations and units with distinct constituents, properties, and
characteristics that greatly affect the design and investigation of a transportation project. For
example:

Many of the volcanic rocks that compose the Coast range, Willamette Valley, and Cascades
can exhibit deeply weathered soil horizons with isolated zones of less weathered materials,
interbeds of weak tuff and other unconsolidated tephra.

Many of the coastal and inland valleys contain deep, soft sedimentary deposits formed by a
rising sea level at the end of the Pleistocene.

The Klamath Terrain in the southwestern portion of the State is a complex mixture of
materials that present difficult conditions for the exploration as well as construction.

Numerous published and unpublished documents are available that provide enough information
upon which to base a background study. Naturally, many portions of the State have more information
than others depending on population densities and previous site uses. However, some basic
information is available throughout the state that can be used for most projects. The following
sections provide a discussion of the most common publications and how they contribute to a
background project study.
Procedures and techniques for the interpretation of maps, aerial photographs, and other remote
sensing products can be found in a wide variety of texts and other publications. Several engineering
geology textbooks provide a good background in geologic interpretation for engineering projects.
However, landform recognition methods are also very well presented in numerous geography texts
and other related books devoted entirely to remote sensing and/or GIS. Geologic interpretation with
specific emphasis on landslides is treated in Chapter 8 of the 1996 TRB Landslides publication.
Topographic Maps
The U.S. Geological Survey (USGS) prepares and publishes 7.5-minute topographic maps at a
scale of 1:24,000 for the entire State, and for most of the rest of the U.S. Topographic maps can be
used to extract both physical and cultural information about the landscape and their consultation
should be the first step in any site investigation. Contour lines provide information about slopes as
well as indications of the underlying geology and geomorphology. The drainage patterns that
develop in the contour lines also suggest geologic and human factors that may have influenced site
conditions. Transportation and development patterns portrayed on USGS quad sheets are an oftenoverlooked source of information. Many roads are aligned to avoid existing geologic hazards or
areas where construction difficulties are expected such as wetlands, steep slopes, or hard, resistant
rock cuts. Quarry and mine site locations are also an important clue with respect to the location and
distribution of bedrock materials.
15-minute topographic maps, also produced by the USGS at a scale of 1:62,500 are also commonly
available, but since they have been discontinued in favor of the 7.5 quad sheets, are becoming
increasingly rare. The advantage of the 15-minute maps is that they can be very old and may show
how land-use has changed in an area since their original survey. Previously existing wetlands that
have since been filled or drained, waste areas, quarries, abandoned mines and other problematic
areas with respect to transportation projects may be identified. Topographic maps should always be
used to identify the arcuate headscarps and hummocky terrain indicative of landslides, wetlands, and
general site accessibility with respect to investigation as well as construction.
Sources of Aerial Photos
Aerial photography is the most common, reliable, easy to use, and usually the cheapest source of
remote sensing available. Aerial photos are very useful in planning subsurface investigation
programs from gaining general knowledge regarding the geology, the extent and distribution of
Volume 1
Manual

ODOT Geotechnical Design


8

April 2010

materials, the location of geologic hazards, potential for encountering contaminants, and determining
access for exploration equipment.
Aerial photographs are widely available through a variety of sources. The ODOT Geometronics Unit
would be the first source for aerial photos as their archives date back to the early 1950s and primarily
cover the areas around the States highways and the Oregon coastline.
Instructions and forms for ordering aerial photographs from the ODOT Geometronics Unit will be
found on the Agencys website at:
http://egov.oregon.gov/ODOT/HWY/GEOMETRONICS/AerialPhoto.shtml.
Additional sources of aerial photography are:
ODOT Geometronics Unit
http://egov.oregon.gov/ODOT/HWY/GEOMETRONICS/AerialPhoto.shtml
The US Geological Survey
http://www.usgs.gov/science/science.php?term=700
http://topozon.com
USGS EROS Data Center
http://edcwww.cr.usgs.gov/
The USDA Aerial Photo Archives
http://www.apfo.usda.gov/
Bureau of Land Management
http://www.blm.gov/nstc/aerial/
University of Oregons Aerial Photography Library
http://libweb.uoregon.edu/map/orephoto/mapresearch.html
WAC Corporation
http://www.waccorp.com/
Spencer B. Gross, Photogrammetric Engineering
http://www.sbgmaps.com/
Intermountain Aerial Survey
http://www.ias-map.com/
Bergman Photographic Services
http://www.bergmanphotographic.com/
Many County Surveyor and/or Assessors offices throughout the State are an additional source of
aerial photography. There are also a number of internet resources for low-resolution images for site
location or other less-detailed applications.

Volume 1
Manual

ODOT Geotechnical Design


9

April 2010

General Use of Arial Photography


Aerial photographs may be taken on either black and white or color film. Each of them have
characteristics that make them superior to one another for different applications although color
photographs are generally considered better since many objects are easier to identify when shown in
their natural colors. Things to consider include:

Color photos also allow for the application of color contrasts and tonal variations to
interpretations. In some circumstances, black and white photographs allow the geologist or
engineer to resolve changes in slope or elevation that may otherwise be lost in the subtle
color changes when using natural color aerial photos.

Another, less commonly available type of aerial photograph are those taken in false color or
infrared (IR). Color IR photography responds to a different electromagnetic spectrum than
natural photography. Differences in soil moisture, vegetation type and soil and rock exposure
are more readily identified on color IR film.

Ideally, both black and white as well as color photos of a site should be analyzed for a
complete analysis of all features unless color IR photos are available in which case it is
generally agreed that for engineering geologic interpretation, natural color and color IR
transparencies provide the best information.

With a general understanding of the site geology, the lateral extent of certain geologic features and
deposits can be estimated from aerial photography. With a stereo-pair of photographs, the vertical
extent can also be estimated in some circumstances. The use of stereo-pairs significantly increases
the ease and accuracy of geomorphic interpretation. Subtle landforms may be discerned that may
otherwise be hidden from view either on-site or on a two-dimensional image.

Geomorphic Identification from Aerial Photography


Landform identification regularly allows the general subsurface conditions to be determined within the
boundaries of that particular feature and thus, an opening impression of the materials to be
encountered. Recognized landforms result from particular geologic mechanisms that allow such
determinations to be made. These landforms are formed by distinct processes such as fluvial,
glacial, or Aeolian and so they are composed of particular materials and compositions. Drainage
patterns that develop within or as a result of certain landforms and geologic structures can be used
as a diagnostic feature when studying aerial photographs. One of the more important landforms to
distinguish during a preliminary study of aerial photographs is landslides. Landslides are readily
identified by their characteristic arcuate headscarps, patterns of disturbed soil and vegetation,
standing water on slopes with no apparent source or discharge (sag ponds), abrupt changes in
slope, disrupted or truncated drainage patterns, and upslope terraces.

Other Applications of Aerial Photography


Vegetation is another important feature to evaluate on aerial photographs since it frequently reveals
certain subsurface conditions. Vegetative cover is related to numerous factors including soil
development on certain bedrock units, depth of the soil profile, drainage and natural moisture
content, climate, and slope angle. The relationship between clear-cutting of forests and debris flows
or adjacent land instability is becoming increasingly important. Consequently, identification of such
conditions within or near a project site is essential. In addition to the geologic characteristics, the
condition or absence of vegetation may be a sign of soil contamination. Zones of dead or discolored
vegetation can indicate the presence of a spill or chemical dump site that field exploration crews may
not be prepared to encounter.
It is also important to review a sequence of aerial photographs from different years to determine the
history of site use and the natural or human-caused changes that have occurred. Significant
Volume 1
Manual

ODOT Geotechnical Design


10

April 2010

changes in the ground contours and shapes can indicate changes due to geologic processes such
as landslides, erosion, and subsidence or changes due to construction on the site such as filling and
excavation. Other aspects of the sites history that can be determined are the activities that occurred
on site such as chemical processing, fuel storage, waste treatment, or similar activities which may
leave contaminated or other deleterious materials behind.

Geologic Maps
The Oregon Department of Geology and Mineral Industries (DOGAMI), USGS, US Department
of Energy, and other agencies publish geologic maps of most of the state at various scales. The
USGS has published a map of the entire state at a 1:500,000 scale. These geologic maps generally
use the USGS topographic maps as a base layer. Geologic maps portray the distribution of geologic
units and provide a general description of each that includes the rock or sediment type, geologic age,
origin, and brief summary of its properties and physical characteristics. Additional information
concerning geologic hazards, groundwater, and economic geology is typically included.
DOGAMI also publishes special studies on geologic hazards in certain heavily populated or
problematic areas of Oregon. Geologic Hazard maps are generally produced to portray specific
themes such as slope stability, liquefaction potential, amplification of peak rock accelerations, and
potential tsunami inundation zones. Such maps provide a general indication of the extent and
magnitude of the hazards they were produced to portray.
Geologic maps for the state are available from DOGAMI and at most of the State Universities
libraries.
Publications are also available for purchase on line from DOGAMI at
http://www.naturenw.org/. In addition, many local agencies and municipalities have contracted for
hazard mapping and planning. These publications may be available from the local agency offices.
DOGAMI is now in the process of a digital map compilation for the state. This compilation allows for
the electronic querying of geologic information published in a selected area. The geologic
information contains pertinent engineering characteristics in many areas. Currently, the compilation
map for the NE sextant of the state is available on CD.

Soil Surveys
The US Department of Agriculture, Soil Conservation Service has published soil surveys for all of the
counties in Oregon. Although these reports are intended for agricultural use, they provide valuable
information on the surficial soils in and around a project area. These bound volumes include maps
and aerial photographs showing the lateral extent of soil units and a description of the overall
physical geography including local relief, drainage, climate, vegetation, and description of each soil
unit together with its genesis. Commonly, the soil units are overlain on a topographic and aerial
photographic base. The reports contain engineering classifications of the surficial soil units, a
discussion of their characteristics such as drainage and susceptibility to erosion suitability for use in
some construction applications.

Remote Sensing and Satellite Imagery


Remote sensing, by the largest definition, involves the collection of data about an area without actual
contact. By this definition, the previously discussed methods of air photo and map interpretation
would be classified as remote sensing. However; for this section, remote sensing is restricted to
imagery obtained by systems other than cameras, or images that are enhanced to distinguish
different characteristics of the earths surface.
Remote sensing as discussed in this section generally utilizes sensors that detect particular
electromagnetic energy spectra that is mostly generated from the sun and subsequently reflected or
emitted from earth. In addition, active systems that transmit and detect energy from the same
platform such as an airplane or satellite are also used to collect imagery. The primary purpose of this
distinction is that aerial photographs allow examination of images in the electromagnetic spectrum
Volume 1
Manual

ODOT Geotechnical Design


11

April 2010

visible to the human eye. Other imagery allows examination of features with reflectance or energy
emission properties that are either outside the spectrum visible to humans or occur with other
features with overlapping spectral reflectance that obscures them to the human eye. Examples of
these other remote sensing systems are; Multispectral Scanning Imagery (MSS), Thermal Infrared
Imagery (Thermal IR), Microwave Imagery (Radar), and Light Detection and Ranging (LiDAR).
Despite their advantages, these remote sensing systems are not a substitute for stereo photographs
and their higher detail, interpretive returns, and overall economy. They are merely a tool to allow
additional interpretation capability for engineering geologic studies.
Thermal Infrared Imagery
These systems obtain images from the thermal wavelength range, generally from 8m to 14m, and
contain the energy emitted from the earth that was previously stored as solar energy. The thermal
properties such as conductivity, specific heat and density of various materials produce different
responses to temperature changes. Such responses can be measured to allow differentiation of
various surface materials. In a sense, thermal IR imagery can be described as a photograph of the
earths albedo.
Obviously, the longer wavelength of thermal IR images will result in a much lower resolution than a
corresponding photographic image. For this reason, thermal data is used to enhance images of
areas with certain surface conditions that are not generally detected by aerial photography. In this
regard, areas composed of materials with similar or overlapping reflectance properties may not show
up on an aerial photograph, but their different thermal properties will make them stand out on a
thermal IR image.
The primary uses of thermal IR imagery are for mapping changes in soil and rock compositions and
anomalous groundwater flow characteristics on an aerial photograph base. Typical engineering
geology applications of thermal IR imagery are:
Fault delineation
Locating seepage at soil and rock contacts
Mapping variations in weathered rock profiles
Mapping near-surface drainage
Multispectral Scanning Imagery (MSS)
MSS systems produce imagery from several distinct ranges, throughout the photographic and
thermal spectrum. These distinct spectra are typically referred to as a band. Each spectral is
concurrently recorded by the scanning instruments along the aircraft or satellite flight line. Much of
the data available came from the Landsat satellite program during the 1970s and 1980s. The early
Landsat satellites used only four spectral bands and achieved a resolution of about 80 meters. Later
satellites used 7-band sensor array with a 30-meter resolution from 6 of those bands. The seventh
was a thermal IR sensor. Special aircraft flights with 24-band sensors can also be obtained.
Images from MSS data can be used to examine the spectral signatures and reflectance of surficial
materials and objects. Different soil and rock materials, as well as the extent of rock weathering, can
be identified by comparing color variations from the different spectral bands. MSS image analysis for
engineering geology is typically used to identify major landforms and tectonic features. Also, the
length of time over which the images were collected allows observation of changes in vegetation,
land use, and the locations of catastrophic events such as fault rupture, flooding, and landslides. As
with thermal IR imaging, MSS is generally used as an enhancement of aerial photography rather
than a substitute for it.
Microwave Imagery (Radar)
Volume 1
Manual

ODOT Geotechnical Design


12

April 2010

Radar utilizes electromagnetic energy from the microwave spectrum, typically with wavelengths from
1mm to 1m. Radar imaging may come from either an active or a passive system. In this regard,
passive systems are a form of thermal IR imaging using the wavelengths that increase to the range
of microwaves whereas active systems emit pulses of energy that are transmitted to the earths
surface where they are reflected back to a receiver.
The most common technique for this type of imagery is Side-Looking-Airborne-Radar (SLAR). For
this technique, the radar scans a portion of the earths surface laterally from an aircraft in a direction
perpendicular to the flight line and at a depression angle measured downward from the horizontal.
Overlapping images created from this method allow stereo viewing of surface features and objects.
Objects that are more perpendicular to the pulse provide a strong energy return to the receiver while
smooth or horizontal surfaces reflect the energy away from the receiver resulting in a dark image. It
then follows that reflection angles and surface roughness as well as vegetation and moisture content
influence the energy returned to the receiver. Objects and features extending above the surface
project radar shadows that are related to the angle of incidence of the energy transmitted and
received. These shadows accentuate the surface topography and thus, structural trends.
SLAR images are typically used in an engineering geology application to identify the surficial
expression of geologic structures, drainage features, structural patterns, and trends. SLAR imagery
is complimentary to aerial photography and should not be a substitute for it. However, SLAR images
have many advantages that provide additional information that is difficult to extract from an aerial
photograph. Their primary advantage is the enhancement of major features that are obscured by the
greater detail of an aerial photograph. Another advantage of SLAR is the ability to obtain clear
images at night and in heavy cloud cover.
Light Detection and Ranging (LiDAR)
This relatively new technology utilizes an active system that is similar to radar in the manner by which
it creates an image. In this regard, energy is emitted from a source and reflected from the earths
surface back to a receiver. However in this case, a laser is used to measure the distance to specific
points and generates a digital elevation model of the earths surface similar to standard
photogrammetric methods. LiDAR equipment is typically mounted in an aircraft although numerous
ground-based applications have been developed that are beneficial to highway engineering geology,
and in particular, rock slope design.
The primary advantage of LiDAR is during post-processing of the data that allows vegetation to be
stripped from the data to provide a bare-earth terrain model. This is a particularly useful technology
in much of Oregon where heavy vegetation obscures much of the ground surface. Landforms that
would typically be obscured stand out in sharp resolution on a LiDAR image where the vegetation
has been removed. In addition to vegetation, structures and dwellings can also be removed. This is
also advantageous where development has occurred over large, ancient structures to the extent
where they completely obscure its features. Disturbed areas and earthworks are also plainly visible
on bare-earth LiDAR images. This allows clear distinctions to be drawn between fills and
embankments, and natural ground surfaces. Bare-earth models also provide a clear resolution of
existing stream courses and channels. Other imagery and photogrammetry-derived mapping often
contain erroneously-located stream segments due to forest cover and/or ongoing lateral migration.
LiDAR images not only provide an unmistakable location of the stream course, but also a clear
rendition of the stream banks and terraces.
ODOT currently stores LiDAR bare-earth and reflective imagery files on the GIS server as hillshade
images and Digital Elevation Models (DEM) files. This server is accessible on the ODOT system and
located at:

Volume 1
Manual

ODOT Geotechnical Design


13

April 2010

\\Sn-salemmill-1\GIS\IMAGES\LIDAR.
Raw ASCII and .LAS-format files are available from ODOTs GIS unit as requested. In order to load
the raw or binary datasets, an external hard drive of at least 500 GB capacity must be provided as
these files are extremely large. LiDAR imagery and DEMs are normally viewed, manipulated, and
analyzed with GIS software and specific GIS software extensions. Specialized software is also
available for LiDAR data and imagery analysis. ASCII and .LAS files can be used to produce a .dtm
file compatible with later versions of Bentley InRoads.
Numerous contractors are available that can provide LiDAR data products; however, ODOT
participates in the Oregon LiDAR Consortium (OLC) for new acquisitions. The Oregon Department
of Geology and Mineral Industries (DOGAMI) was given a legislative mandate to extend LiDAR
coverage throughout the state. The consortium model was approved for funding, collection, and
sharing new LiDAR datasets. DOGAMI, as head of the consortium retains the LiDAR contractor and
develops cooperative agreements between consortium members. The consortium benefits all
members by provided additional coverage for lower cost. As the aerial extent of each acquisition
order increases, the cost per square mile decreases. In addition to lowering the unit cost, more
contiguous areas of LiDAR data are acquired providing greater benefit to all members. Members of
the OLC include Federal, State, and Local agencies, Tribal governments, private entities, and not-forprofit organizations.

2.2.5

Site Reconnaissance

2.2.5.1 General
The purpose of site reconnaissance in geotechnical project planning is to verify the results of the
office study completed in Section 2.2.1 and Section 2.2.2, and to begin formulation of a site-specific
exploration program that will address the issues identified, and determine some of the logistics
required to complete the next phase of investigation. At this stage, the geotechnical designer should
know what to look for at the site, and, with preliminary or conceptual plans in hand, should observe
the anticipated conditions with respect to the proposed project features. Surficial expression of
features and landforms should be checked on the project plans as well as delineating additional
features noted during the site reconnaissance. It is also important to assure that the project maps
are accurate with respect to the actual site conditions, and that significant features were not
overlooked or misrepresented on the preliminary or conceptual design phase maps. The scope of
the site reconnaissance depends greatly on the site conditions, accessibility, and project complexity.
The value of the site reconnaissance is realized later on in the project through a more efficient and
thorough site exploration and geotechnical design. Therefore; site reconnaissance should be
complete and systematic to achieve the final objectives of the office investigation, and may involve a
significant level of effort in the field depending on the project site itself.

2 . 2 . 5 . 2 Ver i f i c a t i o n o f O f f i ce S t u d y a n d S i t e O b s e r va t i o n s
The topography and geomorphology of a site should be reconciled in the field with what was
anticipated in the office study and shown on any maps or aerial photographs. Review and assess
the following:

Outcroppings, road cuts, stream beds, and any other subsurface exposures should be noted
to verify the anticipated conditions based on the published geologic maps and literature. The
presence of artificial fills should be noted and described with respect to its composition, lateral
extent, and estimated volume.

Volume 1
Manual

ODOT Geotechnical Design


14

April 2010

Surface waters, springs, wetlands and other potentially sensitive areas that may impact the
project work should also be noted. In addition, an effort should be made to identify the 2-year
flood zone for future reference.

Boulders, blocks, and oversized materials in stream beds, or projecting from embankments
should be noted as they may be indicative of obstructions in the subsurface. Such
obstructions are one of the most common sources of changing site conditions claims on
projects that involve pile driving, shaft/tieback/soil nail drilling, and excavations. Oversized
materials observed on the surface may not be encountered during exploratory drilling and
thus, the field reconnaissance may be the only record of their occurrence. In addition to
boulders and blocks, existing, abandoned structures such as foundations and utility vaults
can also be an obstruction to foundation installation and excavation.

Any landslide features observed in the office study should be examined in addition to any
new features discovered during the site reconnaissance. All indicators of unstable slopes
such as springs sag ponds, bent tree trunks, disturbed plant communities, abrupt vegetation
changes, and hummocky terrain should also be noted. Measurement and delineation of all
features and indications of slope stability should be completed during the reconnaissance.
Complete investigation of slope stability affecting a project area necessarily involves areas
that may extend a substantial distance away from the proposed alignment.

The performance of existing and nearby structures should be evaluated during the site
reconnaissance. Evidence of settlement, deformation, tilting, or lateral movement can
indicate site conditions that possibly will impact the project design and further exacerbate the
performance issues during construction.

At bridge sites, the existing footings should be evaluated with respect to stream scour.
Exposed pile caps or footings as well as riprap protection generally indicate that scour has
been a concern at the site previously.

2.2.5.3 Preparation for Site Exploration


Potential boring locations should be identified with respect to the preliminary or conceptual plans
available at the time of the site reconnaissance. Once the locations are determined, an assessment
can be made in connection with how they will be accessed by exploration equipment and personnel.
Many projects can be investigated by routine methods with common equipment. However, for some
projects, site access can cost almost as much if not more than the actual subsurface exploration itself
in many circumstances. Physical site access, traffic control, environmental protection, and many
other issues can arise that increase the complexity, and subsequently, the cost of the exploration
program. Every site is different, so each must be assessed individually to determine what methods,
procedures, equipment, and subcontractors will be needed. Some of the most common issues that
need to be addressed are:

Traffic Control Flagging, lane restrictions, and pilot cars are required when working in or
near the travel lanes. In such instances, traffic will need to be controlled for the entire time
the exploration crew is on site. In other areas, traffic control may be needed while loading or
unloading equipment and supplies. In many areas, lane restrictions are only allowed for
nighttime operations. In every case, all efforts will be made to minimize the impact to the
traveling public.

Equipment Required Determining whether the site can be accessed using a standard
truck-mounted drill rig or whether a track-mounted drill will be needed. It may also be
necessary to consider difficult-access equipment that must be transported by crane,
helicopter, or hand-carried.

Volume 1
Manual

ODOT Geotechnical Design


15

April 2010

Physical Access Considering additional equipment to access a site and analyzing the
cost-benefits of their use vs. other drilling equipment and investigative methods. For some
sites, bulldozers and excavators may be needed to construct an access road for drilling
equipment, barges may be needed for in-water work, and special low-clearance equipment
may be needed for work in and around utilities. Where access roads are problematic due to
environmentally sensitive areas that need to be avoided, overall impact, cost, and
reclamation requirements; alternative equipment or methods should be looked upon as a
potential cost or problem-saving measure where the integrity of the exploration information is
not compromised.
For in-stream work, project scheduling becomes a significant issue since restrictions will
be imposed on the times of the year when such activities will be allowed. Furthermore,
the logistics of carrying out in-water work bring additional requirements such as
determining the draft of the barge needed for the depth of the water, how the barge will
be anchored, where the barge will be launched from, how the crew will access the barge
during a shift change, and determining the effects of tidal or current changes on the
drilling operations. A marine surveyor should be engaged for particularly complex overwater operations, and on some waterways, their review of operations is required.
Where bridges are replaced at their present location, and conditions allow, drilling may
be conducted through the existing bridge deck although efforts must be made to assure
that only the deck and not the superstructure are penetrated.

Drilling Conditions Where high groundwater levels, deep water, and loose or heaving
sands and gravels, and obstructions are anticipated, the appropriate drilling methods and
materials should be specified.

Materials and Support Remote locations may require special considerations for
supporting the field crew and the equipment. In this regard, additional logistics may be
needed for delivering drilling supplies, fuel, lubricants, etc., and for the timely delivery of
samples back to the laboratory and office. All-terrain vehicles may be needed to support the
drill crews in such situations, or else preplanning needs to be carried out to schedule or
arrange for extra site provision. Locations for drill water should be identified ahead of time,
and where an ODOT facility is not available, permits will need to be obtained ahead of time
for fire hydrants, private sources, or extraction from streams and lakes.

Right-of-Way The methods by which permits of entry for exploration on private property are
obtained vary from region to region, and frequently, within a region. For all cases, the region
Right-of-Way section in which the project is taking place should be consulted prior to
exploration, and then notified in advance, when and which private properties will be
accessed. The Right-of-Way section manager or their subordinate will recommend either a
standard permit of entry form, or they will obtain the permit of entry internally.
In many instances, private property owners will refuse to grant entry. For these, the
right-of-way section will be required to handle the negotiations for site access, and
determine the terms and conditions.

Utility Conflicts During the site visit, the location and type of utilities should be noted. The
names and contact information located on the utility risers, stakes, and poles should be
recorded. In all cases, the Utility Notification (One-Call) Center must be contacted at least 2
working days prior to commencement of site operations at 1-800-332-2344. The One-Call
Center will recount the utility services that they will notify based on their records. The
geotechnical designer or drilling supervisor will be responsible for notifying any other utilities
operating in the area based on their observations of facilities during the site reconnaissance.

Volume 1
Manual

ODOT Geotechnical Design


16

April 2010

Responsibility for maintaining the utility location markings during site operations belongs to
the field exploration crew.

2 . 2 . 5 . 4 R e c o n n a i s s a n c e Do c u m e n t a t i o n
During the field reconnaissance, photographs should be taken of all the predominant features
previously discussed. Each photograph should be appropriately labeled with the object of the photo,
the direction it was taken, where it was taken from, the date, and ideally, the latitude and longitude of
the photographs origin obtained with GPS equipment.
The observations taken during the site visit should be documented in a memorandum or short
reconnaissance report depending on the scope and complexity of the project. The report should
provide a list and a description of all the observations made, and the prominent features encountered
during the office study and site reconnaissance. Each feature should be located with reference to the
project stationing or reference grid. Once again, there is considerable benefit to locating features
with GPS equipment for long-term record keeping. Project stationing can change, projects can be
postponed for long periods of time, and future projects will occur that will utilize this document see
Section 2.2.1.1. Preplanning for geotechnical design is correlative to any other investment; the
earlier in the process the work takes place, the longer the benefits can be reaped.

References

American Association of State Highway and Transportation Officials, Inc., 1988, Manual on
Subsurface Investigations.
C.H. Dowding, Ed., Site Characterization & Exploration, ASCE Specialty Workshop Proceedings,
Northwestern University, 1978.
U.S. Department of Transportation, Federal Highway Administration Evaluation of Soil and Rock
Properties, Geotechnical Engineering Circular No. 5, FHWA-IF-02-034, April, 2002.
Turner, Keith A., and Schuster, Robert L., Eds., LANDSLIDES Investigation and Mitigation,
Transportation Research Board Special Report 247, 1996, Pages 140-163.
Geotechnical Investigations, U.S. Army Corps of Engineers Engineering and Design Manual, EM
1110-1-1804, January 2001.

Volume 1
Manual

ODOT Geotechnical Design


17

April 2010

APPENDIX 2-A Geology / Geotechnical QC


MATRIX
Table 2-1. Geology / Geotechnical Matrix Checklist QC Check #1 Scoping
Geology
YES

NO

Geotech
N/
A

YES

NO

Rock Slopes
N/
A

YES

NO

N/A

Scope
Project Name and Key Number
Existing structures, earthworks and known
hazards
Proposed structures and earthworks
Design Narrative, defined project area
Project Geography
Bodies of water
Terrain Features
Climate
Region
Project Geology
Province
Bedrock and Quaternary Geology
Structural Geology
Geologic Hazards
Geomorphology
Geologic Impacts/Performance of existing
structures
Performance of existing structures
Previous design efforts in the project area
Cost Estimates for Proposed Work (Design and
Construction)
Monitoring period
Summary of findings and project implications

Volume 1
Manual

ODOT Geotechnical Design


18

April 2010

Table 2-2. Geology / Geotechnical Matrix Checklist QC Check #2 Scope of Work


Geology
YES

NO

Geotech
N/
A

YES

NO

Rock Slopes
N/
A

YES

NO

N/A

Project Scope
Schedule of work
Geology Scope of Work
Geotechnical Scope of Work
Rock Slope Scope of Work
Exploration Scope of Work
Geology project budget
Geotechnical project budget
Rock slopes project budget
Monitoring period schedule and budget

Volume 1
Manual

ODOT Geotechnical Design


19

April 2010

Table 2-3. Geology / Geotechnical Matrix Checklist QC Check # 3 EIS


Geology
YES

NO

Geotech
N/
A

YES

NO

Rock Slopes
N/
A

YES

NO

N/A

Survey of proposed alignments and alternatives


Bedrock units to be encountered
Surficial units to be encountered
Physical geography effects on proposed
alignments and/or slope geometries
Location
Extent
Climate
Topography
Geologic Province
Character of expected geologic units and their
performance history
Geologic hazard potential
Summary of known geologic hazards
Summary of known geologic impacts to existing
features
Performance of structures and earthworks
along proposed corridors or alignments
Known
geotechnical-related
problems
in
existing structures and earthworks in the
proposed project area
Mitigation methods and costs for potential
geotechnical issues
Geotechnical
characterization/estimated
properties of geologic units
Discussion of the performance of project area
materials and geologic units
Correlation of properties of expected materials
with similar studies
Cost-benefit analysis of proposed alignments
and/or locations

Volume 1
Manual

ODOT Geotechnical Design


20

April 2010

Table 2-4. Geology / Geotechnical Matrix Checklist QC Check # 4 Concept


Geology
YES

NO

Geotech
N/
A

YES

NO

Rock Slopes
N/
A

YES

NO

N/A

Concept Plan Review


Reconnaissance Report (File Summary Survey)
Consultation of published literature
Consultation of unpublished literature
Aerial photographs and
other remote sensing
Aerial photographs from
different years to review
varying conditions through
time and site history
As-built plans
Maintenance records
Region file survey
Consultant reports
RHRS/Unstable slope
inventory
Review of maintenance activities that have
affected the site (e.g. rockfall containment,
slope stability, drainage)
Review of geographic and geologic conditions
affecting slope stability with respect to
conceptual evaluation of landslide/rockfall
remediation schemes
Determine the potential affect of outside
stakeholders on the remediation options (USFS,
Gorge Commission, Tribal Governments, etc.)

Volume 1
Manual

ODOT Geotechnical Design


21

April 2010

Table 2-5. Geology / Geotechnical Matrix Checklist QC Check #5 Exploration Plan


(10% TS&L)
Geology
YES

NO

Geotech
N/
A

YES

NO

Rock Slopes
N/
A

YES

NO

N/A

Exploration Plan
Exploration Plan Summary
Survey Requirements
Work Products
Scope, Schedule, Budget
Project Features requiring subsurface
investigation
AASHTO compliance for project features
Boring/Exploration spacing
Boring/Exploration depth
Sampling frequency
FHWA recommended standard practices for
rock slopes
Evaluation/inclusion of alternative or
supplementary exploration methods
Consideration of alternative tests and/or
techniques that would provide better quality
and economy
Appropriate rock slope mapping and drilling
programs for the proposed mitigation measure
Evaluation of the expected site conditions and
compatibility
with
standard
exploration
procedures
Minimum explorations for trenchless
installation and associated features

pipe

Exploration Plan Review


Structures
exploration

and

earthworks

for

Proposed exploration at each structure


location

Volume 1
Manual

ODOT Geotechnical Design


22

April 2010

Table 2-6. Geology / Geotechnical Matrix Checklist QC Check #6 2/3 TS&L)


Geology
YES NO

N/
A

Geotech
YES NO

N/
A

Rock Slopes
YES NO N/A

Field Exploration Review


Site-specific field explorations
Borings
Test Pits
Hand-auger holes
Geophysics
In-Situ testing
Site and vicinity reconnaissance
Project-level geologic mapping
ASTM conformance
Drilling methods
Sampling and testing
Deviations from standards noted
and described
Review of alternative tests or techniques
Quantity of samples for laboratory testing (collection
and recovery)
Adequate samples and laboratory testing to
characterize and determine the extent of subsurface
materials
Undisturbed
samples
in
cohesive
and/or
compressible materials
Core drilling procedures
ODOT standard core box placement
and labeling
HQ or larger-sized core diameter
Triple-tube recovery system
Recovery appropriate for the materials
encountered (never less than 80%
unless special conditions exist)
Core specimens labeled and
photographed while wetted
Legible and appropriate core
photography
Specimens removed for laboratory
testing replaced in the core box
with the appropriate marker
Drilling techniques correspond to the materials
encountered
Augers used while investigating for the piezometric
surface in soil
Indication
where
natural
moisture
content was altered by introduced
fluids

Volume 1
Manual

ODOT Geotechnical Design


23

April 2010

QC Check #6 ( TS&L) (continued)


Geology
YES

NO

Geotech
N/A

YES

NO

Rock Slopes
N/A

YES

NO

N/A

Methods used to determine piezometric surface in rock


Fluids used to stabilize boreholes in sandy material or other
heaving conditions
Measures to avoid affecting SPT and other
values and intervals in
heaving conditions

testing

Drilling activities recorded on standard boring log forms


Fluid return and color changes
Drill action and rate
Shift/personnel changes
Bit wear
Drilling techniques
All information used for interpretation
subsurface conditions
Locations where groundwater was

of
encountered

Open hole water levels recorded


beginning of each drilling shift

at

the

backfill

and

Dry holes specifically noted


Types, quantities,
sealing materials

ad

depths

of

Soil and rock materials identified,


classified,
and
described according to
the current version of the ODOT
Soil
and Rock Classification Manual
Complete soil and rock descriptions
Additional physical properties,
diagnostic,
distinguishing features
recorded on the logs
Boring locations surveyed with
Coordinates and true elevations

respect

to

State

or
Plane

Conversion to SPC/true elevation where assumed values are


used
Borings referenced by project stationing
Borings referenced by bearing and distance to permanent
features or reference points in the absence of an existing base
map or survey
Preliminary subsurface drawings and/or model for adjusting
exploration according to current findings

Volume 1
Manual

ODOT Geotechnical Design


24

April 2010

QC Check #6 ( TS&L) (continued)


Geology
YES

NO

Geotech
N/A

YES

NO

Rock Slopes
N/A

YES

NO

N/A

Boreholes abandoned according to Water


Resources standards
Instruments installed according to their
purpose (e.g. inclinometers installed below
the slide plane, piezometer sensing zones in
the water-bearing strata, etc.)
Records of piezometer casing type/size,
slotted zones, slot size/frequency
Records of sealing and filter pack placement,
sizes and grades of the materials
VWP Installations
Manufacturers calibration sheets
Field calibration results
Initial reading
manual observation

consistent

with

Inclinometers
Appropriate slurry mixture
Slurry quantity recorded
Distinct zones of grout-take noted
A0 direction noted, proper A0
inclinometer alignment
Tube stick-up recorded
Water Resources Hole Reports completed
correctly and filed within the 30-day
requirement
Appropriate rock mass classification system
used to evaluate rock slope excavation
performance
Rock slope surface mapping
Overburden thickness and type
Discontinuity
thickness,
type,
surface roughness, spacing, orientation, and
shape
the

Zones of differential weathering on


slope
Location and volume of seeps and
springs

Volume 1
Manual

ODOT Geotechnical Design


25

April 2010

QC Check #6 ( TS&L) (continued)

Volume 1
Manual

ODOT Geotechnical Design


26

April 2010

Geology
YES

Geotech
NO

N/A

YES

Rock Slopes
NO

N/A

YES

NO

N/A

Preliminary Geotechnical Recommendations


TS&L Foundation Design Memo
Description of proposed project
Anticipated subsurface conditions
Preliminary
foundation
recommendations

design

Foundation types
Preliminary capacities
Rational for selecting the
recommended foundation
type and capacity
Discussion of liquefaction potential and
associated effects
Suggested retaining wall types
Preliminary slope recommendations
Site Model Review
All exploration locations located on plan view
maps referenced to the project
Plan view maps developed to the appropriate scale
to show the necessary features with respect to the
overall project
Appropriate plan map contour interval
and labeling
Borehole collar elevations consistent
with nearest contours
Standard map elements
Cross-sections, fence diagrams, profiles and/or
block diagrams used to display the 3-dimensional
distribution of geologic units, features, structures,
and engineering properties
Geologic model consistent
properties of defined units
Material properties/laboratory
recorded on the drill logs

with

engineering

testing

results

Laboratory testing used to develop engineering


geologic units
Laboratory testing results displayed graphically to
support the engineering geologic model (e.g.
graphs or charts plotting engineering properties
with depth or along a graphic lithology column)
Laboratory testing program included samples from
each boring or test pit to confirm the field and
visual classification
Laboratory results incorporated into the final drill
logs and subsurface model
Laboratory testing
to
verify or
confirm
interpretations or further characterize a unit

Volume 1
Manual

ODOT Geotechnical Design


27

April 2010

QC Check #6 ( TS&L) (continued)


Geology
YES

Geotech
NO

N/A

YES

Rock Slopes
NO

N/A

YES

NO

N/A

Final drill logs match the interpretive drawings and


preliminary drawings for the Geotechnical or
Foundation Datasheets
Clear distinction between observed and inferred
features and relationships in the geologic model
Review laboratory test results to determine if
modifications are required in specific geologic
units at different locations in the subsurface model
Process developed to incorporate laboratory
testing to assure correct and consistent material
classification and description between borings and
to develop engineering geologic stratigraphy from
the test results
Review physical properties testing to determine if
initially misidentified materials occur elsewhere in
the project subsurface
Related soil classifications modified as a result of
physical properties test results
Results of instrumentation programs match the
engineering geologic model
Geologic model encompasses the project design
details to show the effect of the geology on the
facility
Proposed cut lines, excavations,
tunnel/pipe alignment, and foundations
all plotted in the subsurface model
Geologic features affecting the design
such as seeps, springs, piezometric
surfaces, and daylighted adverse
structures clearly shown and identified
in the model
Blocky or rubble-zones that could
produce overbreak in rock cuts or
excavations
Boulders or other obstructions in
proposed excavations or pile and shaft
foundations
Groundwater surfaces
Delineation of collapsible or expansive
soils
Cuts or fills on known or potential slide
areas

Volume 1
Manual

ODOT Geotechnical Design


28

April 2010

QC Check #6 ( TS&L) (continued)


Geology
YES

Geotech
NO

N/A

YES

Rock Slopes
NO

N/A

YES

NO

N/A

Foundations in or near bog/marsh


areas
Excavations below the groundwater
surface, determination of the amount of
water that will be encountered and
the effect of
piezometric drawdown on groundwater
resources
Delineation of potentially soft subgrade
on the project plan map
Geologic
interpretation
of
materials
and
stratigraphy
incorporates
the
engineering
properties of the strata encountered (e.g. geologic
units are subdivided down to the level of distinct
engineering properties)
Cross-cutting relationships established
Quaternary-aged
identified

features

and

discontinuities

Determine if weak or weathered rock sources


identified for use on the project are likely to be
friable or nondurable
Slake Durability testing of exposed rock face
material
Thorough representation of materials tested for
strength and compressibility rather than reliance
on empirical correlations, especially those based
upon Standard Penetration Tests
Appropriate
strength
tests
distinguish between drained
conditions where needed

conducted
to
and undrained

Determine if the total stress envelope of the CIU


test with pore pressure measurements has been
used improperly to define the relationship of
undrained shear strength with depth
Determine if the existing and proposed state of
stress has been accounted for during strength
testing
Evaluation of consolidation tests: reconciliation of
the test-derived preconsolidation pressure with the
actual stress history of the sample

Volume 1
Manual

ODOT Geotechnical Design


29

April 2010

Table 2-7. Geology / Geotechnical Matrix Checklist QC Check #7 Preliminary Plans


Geology
YES

NO

Geotech
N/A

YES

NO

Rock Slopes
N/A

YES

NO

N/A

Engineering Geology Report


Geotechnical Report
Rock Slope Report
Preliminary Geotechnical Datasheets
Datasheets completed for all required structures or features
Profiles drawn along project alignment centerlines or specific
offsets
Cross-sections, additional profiles completed
structure-specific information, or to provide
information in areas of complex geology

to show
additional

Sample and property data


Subsurface model used to develop the Geotechnical Datasheets
Subsurface information shown on the datasheets matches the
final logs
Drawings made at appropriate scales to show the needed level
of detail
Interpretation shown on the datasheets
Geotechnical Datasheets completed according to Subsurface
Information Policy
Detail Drawings and Plans
Review geotechnical items in the bid schedule
Assure specification writers review of geotechnical items in the
special provisions
Review specification writers modifications of geotechnical
items in the special provisions
Correct length and locations for buttresses, surface and
subsurface water collection and discharge features shown on
the plans
Correct materials called out on the plans
Sequence of construction for buttresses

Volume 1
Manual

ODOT Geotechnical Design


30

April 2010

QC Check #7 Preliminary Plans (continued)


Geology
YES

Geotech
NO

N/A

YES

Rock Slopes
NO

N/A

YES

NO

N/A

Staged construction sequence for surcharging, wick


drains, and ground improvement
Appropriate drainage discharge locations
Recontouring of slide areas clearly shown
Surface water drainage in slide areas addressed in
the plans or detail drawings
Buttress, drainage, or other features shown with the
correct elevations and dimensions
Slope protection mat and rockfall protection fences
Mesh type
Anchor spacing
Quantities
Special provisions, including those for
high-impact fences
Standard Drawings included in the

plans

Special access issues and requirements


Standard drawings and special
provisions for PVC-coated mesh
Rock Bolts and Dowels
Design Loads
Design Lengths
Locations
Quantities
Corrosion protection
Performance and proof-testing
requirements
Reference to the Qualified Products

List

Rockfall Retaining Structures


Type, Size, and Location
Quantities
Slopes (Rockfall Protection Berms)
Backfill type specifications
Special Provisions
Rock Slope Drainage
Location
Drain lengths
Drain angles and orientations
Quantities
Water collection and disposal

Volume 1
Manual

ODOT Geotechnical Design


31

April 2010

QC Check #7 Preliminary Plans (continued)


Geology
YES

Geotech
NO

N/A

YES

Rock Slopes
NO

N/A

YES

NO

N/A

Shotcrete
Locations
Areas of coverage
Quantities
Anchorage
Reinforcement
Standard drawings and details
Drainage
Performance requirements
Installation details
Temporary Rockfall Protection
Review type for suitability
Locations
Length
Height
Required materials and quantity
Details
Rock Blasting and Rock Excavation
Quantity of Controlled Blast Holes
Overburden slopes and slope breaks
shown on the plans
Special Provisions
Blast Consultants
Noise/vibration monitoring
Preblast survey
Blasting plan review

Volume 1
Manual

ODOT Geotechnical Design


32

April 2010

Table 2-8. Geology / Geotechnical Matrix Checklist QC Check #8 Advanced Plans


Geology
YES

NO

Geotech
N/A

YES

NO

Rock Slopes
N/A

YES

NO

N/A

Preliminary Wall Drawings


Review subsurface information on Geotechnical Datasheets for
retaining structures
Retaining Wall Drawing Review
Type, Size, Location, Height,

Backslope

Quantities
Backfill types
Wall drainage
Special Provisions
Design Changes and Addenda
Design calculations for added structures and features
Design calculations for structures and features that have moved
Review design assumptions
Changed Criteria
Changed Type, Size, Location
Changed Quantities
Additional exploration requirements for added structures or
features
Appropriate exploration carried out for added structures or
features
New data incorporated into the overall geologic interpretation
Further characterization of geologic units with additional data
Resolution or confirmation of previous inferences and
interpretation
Additional risk assessment

Volume 1
Manual

ODOT Geotechnical Design


33

April 2010

Table 2-9. Geology / Geotechnical Matrix Checklist QC Check #9 Final Plans


Geology
YES

NO

Geotech
N/A

YES

NO

Rock Slopes
N/A

YES

NO

N/A

Final Plan Review


Geotechnical or Foundation Datasheets completed for all
structures, facilities, ad features for which they are
required
Geotechnical Datasheets completed
Subsurface Information Policy

according

to

Engineer or Geologist has stamped all sheets that they


are responsible for
Information provided on the datasheets exactly matches
what is presented on the final logs and in the
Engineering Geology report
Final review of detail and plan sheets
Final review of bid item quantities
Final review of Special Provisions

Volume 1
Manual

ODOT Geotechnical Design


34

April 2010

Chapter

Field Investigation

Introduction

For any transportation project that has components supported on or in the earth, there is a need for
subsurface information and geotechnical data during its planning, design, and construction phases.
Any geologic feature that affects the design and construction phase of a project, or has a bearing on
site or corridor selection in terms of hazards and/or economics must be investigated and analyzed.
Of equal importance is the clear and accurate portrayal of these conditions in a format that is
accessible and understandable by all users.
Consider the following during field investigation:

Subsurface investigation: The objectives of a subsurface investigation are the provision of


general information on the subsurface conditions of soil, rock and water, and specific
information concerning the soil and rock properties that are necessary for the project
geotechnical design and construction.

Scale of investigation: For transportation projects in Oregon, the appropriate scale of


investigation must be carefully considered. Because of Oregons geology and geography,
subsurface conditions are complex and may vary widely over short distances. A more
thorough investigation will provide additional information that will generally decrease the
probability of encountering unforeseen conditions during construction, and increase the
quality and economy of the geotechnical design of a project.

Balance of investigation: Time and fiscal considerations will constrain the scale and
resolution of the field investigation. Therefore; the geotechnical designer must balance the
exploration costs with the information required and the acceptable risks.

The technical decisions and details required for site investigations require the input of trained and
experienced professionals. Every site has its own particular circumstances, and diverse geologic
conditions, professional experience, available equipment, and the previously described time and
budgetary restraints all contribute to the most cost-effective site investigations. The implications of
site-specific geologic conditions for the type of proposed facility must be investigated for each project.
The remainder of this chapter describes established ODOT criteria to be used in field investigations
as well as information on any areas where ODOTs criteria differs from the FHWA and AASHTO
guidelines. More information can also be found in the Federal Highway Administration Subsurface
Investigations - Geotechnical Site Characterization Reference Manual (FHWA NHI-01-031).

Volume 1
Manual

ODOT Geotechnical Design


1

April 2010

3.1.1

Established Investigation Criteria

Professional experience and judgment are the basis of any field investigation program. This chapter
is not intended to provide a prescriptive approach to field investigation, however; there are some
established base levels of investigation for transportation facilities that must be mandated to assure
consistency and quality throughout the agency, and to address a common level of risk acceptance.

These baselines were based on Federal guidance and the AASHTO Manual on Subsurface
Investigations, 1988. ODOT has adopted the baseline requirements for subsurface
investigations from the AASHTO Manual.

However, due to the more variable conditions found in Oregon, ODOTs practice is slightly
more rigorous with respect to exploration spacing and sampling. ODOT variance from
AASHTO guidelines is outlined in Section 3.5 and Section 3.6. LRFD Bridge Design
Specifications, Section 10 provides an additional resource for subsurface investigations,
supplementary to the AASHTO guidelines.

The most important component of subsurface investigation is the personnel that direct the field
activities, interpret the information, and present the results in a clear manner to those responsible for
the final geotechnical design and construction of the project. The quality of information produced
from a subsurface investigation can vary substantially depending on the experience and competence
of the personnel charged with its conduct. Radically different interpretations and conclusions can
result from substandard investigation programs. Subsurface investigation is an investment in the
success of a project with returns that range from 10 to 15 times the cost of the investigation later
realized during final design and construction.

General Subsurface Investigation

For most projects, the main purpose of a subsurface investigation program is to obtain the
engineering properties of the soil and rock units and define their vertical and lateral extent with
respect to thickness, position in the stratigraphic column their depth, and aerial extent where they
could affect the design and performance of a structural or earthwork feature.
The properties normally evaluated include Index Properties such as:

natural moisture content, and

Atterberg Limits.
Additional physical properties may be evaluated, such as

shear strength,

density,

compressibility, and

in some cases, permeability.

The location and nature of groundwater is evaluated in every subsurface investigation. In addition to
material properties, subsurface investigations are carried out to explore and monitor geologic
hazards that were identified in the office studies previously conducted.

Volume 1
Manual

ODOT Geotechnical Design


2

April 2010

For this later purpose, landslides are the most common hazard although caverns, compressible
materials, high groundwater, faults, and obstructions may also form the basis or extension of a
subsurface investigation program.

3.2.1

Subsurface Investigations Phases

Subsurface investigations may be carried out with varying levels of intensity depending on the phase
of the project for which they are conducted. The typical phases are described in the following
sections.

3.2.1.1 Phase 1
For the Field Survey and/or Alternative Design phases (Usually described as Phase 1) of a project,
the information gathered from the office study is usually sufficient for preliminary
geologic/geotechnical input to the project team and for completion of the Soils and Geology chapter
of the Environmental Impact Statement (EIS). In the case of a large and/or complex project, or if
geologic conditions will have a major impact on the design and construction of a project, then some
amount of subsurface investigation will be warranted to determine the exact location and extent of
the problems and to devise some preliminary cost estimates and alternatives. Ideally, when
performing a subsurface investigation during Phase 1, the exploration would be situated at the
location of a major project feature that would be investigated later during project design. However, as
this occurs early in the project, or certain other alternatives are under consideration, the precise
locations of bridge bents and final alignments may not be known.

3.2.1.2 Phase 2
The project design phase (Field Survey up to Preliminary Plans, usually referred to as Phase 2) is
where the most intense and focused subsurface investigation occurs for specific project features.
Wherever possible, the project design or Phase 2 investigation should capitalize on any previous
explorations in the project area. Personnel responsible for the field investigation and geotechnical
design should determine the utility of this information.
The project design phase subsurface exploration and testing program provides the geotechnical data
specifically required by the projects geotechnical design team. The investigation provides the
aforementioned informational needs for the foundation and earthworks design as well as:

Additional information applicable to other related project elements such as the chemical
properties of soil with respect to corrosion of structural elements, and issues associated with
environmental protection and erosion control.

The project geotechnical design analyses, decisions, and recommendations for construction
will be based on the information gathered during the Phase 2 investigations.
For these reasons, the information gathered during this phase of investigation should achieve a
degree of accuracy, thoroughness of coverage, and relevancy to support the project design decisions
and to allow for realistically accurate estimates of geotechnical bid items.

Volume 1
Manual

ODOT Geotechnical Design


3

April 2010

3 . 2 . 1 . 3 O t h e r Ph a s e s
There will be some instances where additional subsurface investigation is necessary during
Advanced Plans, Final Plans, or even during the construction phase of a project. This is not
necessarily due to an incomplete investigation during the project design phase, but rather the result
of unforeseeable problems that arise during construction, or late design changes following the main
investigational effort and/or geotechnical design. Subsurface investigation is conducted to provide
design information and is usually adequate, in most cases, for contractors estimates for construction
and bidding. Explorations conducted during construction are uncommon, and are usually carried out
to resolve problems or answer questions that arise while the project is being built.
Occasionally, explorations will occur as part of the construction activity to install and monitor needed
instrumentation. When design changes occur late in a project, additional subsurface investigation
can be necessary to confirm the geotechnical design assumptions or to develop additional
information.

Exploration Plan Development

The Exploration Plan is a document that describes the subsurface investigation activities that will take
place to obtain the engineering properties required for geotechnical design. The objective of the
Exploration Plan is to:

assure that the sampling and testing carried out for the subsurface investigation thoroughly
covers each of the geologic units applicable to the geotechnical design

verify that the maximum amount of information can be obtained from the fewest number of
borings or other higher-cost methods

In order to achieve this, the plan must be updated and modified as exploration proceeds to make
sure that the number of samples taken, and tests performed in each unit provides enough numeric
measurements of each critical engineering property distributed throughout the geologic unit to
provide enough confidence in the property to base the geotechnical design upon. In this regard, the
properties of a material at one end of a long alignment may not hold true for the other end, and a
geotechnical designer will not want to base all design parameters for that material on only one or a
few samples.
Subsurface investigation conducted during the project design phase must fully define the subsurface
conditions at a project site to meet the requirements of geotechnical design and construction. The
proper execution of the Exploration Plan will assure that samples and tests are numerically adequate
and distributed vertically and laterally throughout each geologic unit, and that every important
geologic unit at the site is discovered and investigated to the maximum feasible extent. The
Exploration Plan will also assure that the site investigation is conducted in accordance with the
standards of practice outlined in the 1988 AASHTO Manual on Subsurface Investigations and
augmented in this manual. These standards are further subject to modification due to the variability
of the site geology, sensitivity to potential changes, and risk or potential impact.
Note:
Exploration Plans should be created, reviewed, and executed by an experienced engineering
geologist or geotechnical engineer.
The geotechnical designer should comprehensively evaluate the various methods and procedures
for subsurface exploration that are currently available to maximize the amount of information
Volume 1
Manual

ODOT Geotechnical Design


4

April 2010

gathered while reducing costs to the extent possible. The most common method for achieving this is
to gain the most information from the fewest number of borings.
Alternatively, various types of exploration methods may be used where practical in lieu of the more
expensive borings to realize those cost savings without compromising the necessary acquisition of
information.

3.3.1

Exploration Plan Considerations

One of the leading issues addressed when developing the Exploration Plan is the overall scale or
intensity and level of effort for the subsurface investigation. To answer these questions, the expected
complexity of the project sites geology must be considered with respect to nature of the proposed
project, and the projects requirements from the subsurface investigation.
In effect, there are some primary factors that will necessitate increasing the Exploration Plan for a
larger-scale subsurface investigation including:

complex site geology

complex site conditions

scale of the project

sensitivity of the facility to variations in site conditions

The subsurface investigation program should be scoped according to these issues rather than from
some baseline requirement. Each exploration should be justifiable in terms of the information
needed from it. Such informational requirements form the basis of the following criteria:

the type of boring

location

depth

types of sampling

sampling interval

These questions can only be answered by the experience, knowledge, and application of
engineering geologic principles by the geotechnical designer. Through careful examination of the
results previously obtained by the office study, and their experiences working in the area, they are the
essential resource for determining the requirements of the subsurface exploration program.

3 . 3 . 1 . 1 M i n i m u m R e q u i r e m e n t s f o r S u b su r f a c e
I n ve s t i g a t i o n s
This does not however, preclude the necessity of established minimum requirements for subsurface
investigations. The base level of investigation has value as an initial approach to a subsurface
investigation and for preliminary cost estimation of exploration activities as well as assuring that some
uniform amount of exploration is accomplished for all geotechnical design. The minimum standards
for subsurface investigations are well-defined in the 1988 AASHTO Manual on Subsurface
Investigations and are broadly accepted in the practice.

Volume 1
Manual

ODOT Geotechnical Design


5

April 2010

Where ODOT Differs from the AASHTO Manual


Where ODOT practice differs from the AASHTO Manual is in the divergence from the minimum
amount of investigation. AASHTO allows for a reduction from the minimal amount of exploration in
areas of predictable geologic conditions and the absence of any geologic hazards. Such conditions
generally do not exist in Oregon and as a rule, prohibit any reduction of the exploration program.
Rather, explorations are added to the program due to the unpredictable nature of the states geology.
Much of the work performed during the preliminary office studies will assist in determining the overall
scale of the subsurface investigation program.
Such added expenditures are always justifiable when additional exploration, testing, and analyses
result in correlative savings on the construction cost and in an overall better geotechnical design.

3 . 3 . 1 . 2 R i s k Tol e r a n c e
Further consideration in the development of the Exploration Plan should be given to developing an
assessment of the risk tolerance of the project to unforeseen subsurface conditions. In this regard,
an assessment of the risks assumed by the constructability and function of the design feature without
the benefit of site-specific subsurface information should be conducted with respect to the potential
for cost overruns during construction and to potential for long term maintenance or increased lifecycle
costs. The cost of an over conservative design resulting from a hedge against unknown subsurface
conditions is another aspect of risk that should also be evaluated. This is where a design is forced to
be based on the worst possible condition known to be present or perceived at a site in order to
prevent failure because the lack of information precludes the assessment of other alternatives.
Generally, an evaluation of the potential risks at a project site occurs as exploration progresses and
the variability of the subsurface is discovered.

3 . 3 . 1 . 3 S t r u c t u r e S e n s i t i vi t y
The sensitivity of a structure or other facility in terms of performance to subsurface variability also
influences the scale of the subsurface investigation. Consider the following in relation to structure
sensitivity:

Where settlement is concerned, structures are much more sensitive whereas embankments
overall are able to tolerate more post-construction deflection not withstanding those sections
adjacent to bridges.

Existing structures adjacent to transportation projects also increase the sensitivity of projects
in the built-up or urban environment. Where construction is to occur adjacent to existing
structures or private buildings, the tolerance for settlement or deflection and even vibration is
essentially eliminated, and correspondingly, the need for subsurface information increases.

Such sensitivity can also extend to environmental, cultural, and archaeological sites where
great efforts will be made to mitigate impacts during construction. For these circumstances,
significant efforts in pre through post-construction monitoring are often required with
instrumentation installed far in advance of contract letting.

Certain types of construction may also be more sensitive to unanticipated subsurface


conditions such as drilled shaft installation where relatively small changes can result in a
sizeable cost increase.

Volume 1
Manual

ODOT Geotechnical Design


6

April 2010

Despite the best efforts and most detailed subsurface investigations, every significant subsurface
condition may not be discovered or fully examined. The objective here is to reduce the risks
accepted to the barest minimum, and to have some understanding of the risks that will remain.

3 . 3 . 1 . 4 S u b s u r f a c e I n ve s t i g a t i o n S t r a t e g y
An important strategy when conducting the subsurface investigation is to complete the most
important explorations first with the idea that the project schedule may change, funding may be
terminated, or some other decisions made that preclude the completion of all the planned borings.
From this standpoint, the important borings are those that:
1. provide information about geologic hazards affecting the project or that require monitoring for
mitigation design,
2. provide the information that the engineer needs to design the most critical structures, and
3. again, those locations that provide the most amount of information for the lowest expenditure.
This approach to the subsurface investigation allows design to proceed in the event of the inevitable
project schedule or other priority shifts that may have a more urgent need for geologic or
geotechnical resources. It is quite common for a planned exploration to be interrupted by the needs
of emergency repair work or other critical-path project, and having these explorations complete first
allows engineers to continue work on a project rather than having to wait for the emergency to pass
before getting the information they need to continue so that the interrupted project doesnt become
an emergency itself.
Note:
We recommend referring to Section 7.4.1 AASHTO which provides additional items to consider in
determining the layout of a project subsurface investigation in addition to prioritization of the
explorations. This bulleted list describes key issues in determining importance and priority of
explorations from locations to structures that they are intended for as well as the use of less or even
more expensive methods for investigation that may be required.

3 . 3 . 1 . 5 S c h e d u l e o f Su b su r f a c e I n ve s t i g at i o n s
Subsurface investigations are ideally completed as early in the project as possible to allow sufficient
time for geotechnical design, quantity estimation, and consideration of alternatives. Clearly, many of
the project features must already be known to some degree before the Exploration Plan can be
formulated. Right-of-way needs must be established to determine cut and fill slope angles and
heights or the need for retaining structures. Even more detailed plans are needed to begin bridge
foundation investigations. Typically, the bridge type, size, and location (commonly referred to as
TS&L) must be known in order to obtain ground-truth information at the precise bent locations.

Completion of Exploration Plan


Because of these informational prerequisites, the Exploration Plan is usually completed soon after
initiation of the structure TS&L phase with a goal for completion set at the 10% of TS&L completion
with respect to its timeline. The target for completion of preliminary geotechnical recommendations is
set at 2/3 TS&L.
In order to meet this date, there will be less than 50% of the TS&L timeline to complete the
subsurface investigation and provide the needed information to the geotechnical designer charged
with making the preliminary recommendations.

Volume 1
Manual

ODOT Geotechnical Design


7

April 2010

Subsurface investigation performed during preliminary phases may be called for at any time prior to
Phase 2, particularly during the EIS phase depending on the size of the project or any other special
requirements. These investigations are intended to develop project geotechnical constraints and/or
to provide general information to assist in alternative route selection, and to address particular
requirements of the EIS rather than to gain site-specific geotechnical design parameters. Preliminary
subsurface investigation typically takes place on an existing state right-of-way readily accessible
areas so there should not be additional time and money spent in acquiring permits of entry, building
access roads and reclaiming sites.

Instrument Monitoring Periods


An additional aspect of the subsurface investigation schedule that also needs to be determined is the
requirement for instrument monitoring periods. These are particularly important as they commonly
extend before and beyond typical project timelines.

Landslides: Projects that involve landslide repair or evaluation are the usual reasons for
broadening timelines as it is critical to monitor landslide movements over periods of time that
include at least one wet season (usually November through April) to assess the nature of the
slide evaluate the relationships between precipitation, groundwater, and slide movement, and
determine the correct slide geometry for stability analysis.

Groundwater: It is also important to monitor groundwater for other construction applications


throughout seasonal fluctuations to help determine actual construction-time conditions.
Grading operations or excavations that would be made in-the-dry during certain times of the
year may occur below the groundwater surface during other months. Every effort must be
made to collect this information regardless of the time of year that exploration is conducted.

Post-construction monitoring: Where post-construction monitoring is necessary, it should


also be identified as early in the Exploration Plan development as possible. Critical structures
in addition to landslides may require such instrumentation for quality assurance in addition to
providing an assessment of long-term performance.

3.3.1.6 Exploration Sites


One of the primary factors affecting the schedule of the subsurface investigation program is providing
access to drill sites. This includes acquiring the necessary permits as well as the actual physical
occupation of the drill site.
Note:
Preliminary borehole location should have taken place during the initial site reconnaissance and
major requirements with respect to accessibility should have been identified at that time. Since
access to certain drill sites requires a significant investment of time, it is necessary to start acquiring
permits of entry, environmental clearances, and engaging contractors to build access roads or bring
additional resources to move the drilling equipment.
The geotechnical designer should clearly indicate the necessary borehole location tolerances to the
field crews to assist in determining site access. When situating a borehole, consider the following:

For some sites, a few extra feet of tolerance available will allow a borehole to be accessed
with standard equipment or with minimal disturbance while at others, considerably greater
efforts will be necessary to place the borehole at the precise location.

Volume 1
Manual

ODOT Geotechnical Design


8

April 2010

Where the location of the exploration is crucial, it may be reasonable to mobilize specialty
drilling equipment.

Several factors contribute to the amount of tolerance allowed for an exploration. Among
these are the phase of the investigation for which the explorations are performed, in this
case, the final design explorations would require the more precise location.

The types of structure, expected subsurface conditions, and surrounding facilities also have
more exacting standards for borehole placement.

A spread footing on rock, or a tieback wall adjacent to and supporting an existing structure
are examples of cases where relatively minor changes in the subsurface conditions have
very serious consequences during construction and would therefore warrant the extra
expenditure to precisely locate the explorations. In this case, the expenditure for mobilizing
special equipment would be far exceeded by orders of magnitude from ensuing claims or
even, litigation.

3 . 3 . 1 . 7 R i g h t - o f - Way a n d P e r m i t s o f E n t r y
Determining the exact boundaries of the States right-of-way during exploration planning is essential
since this demarcation is very commonly not correlative to the highway centerline nor does it fall at a
constant length perpendicular to it. Current right-of-way maps should be consulted to assure the
correct property ownership at the exploration site or for any land that must be traversed by
exploration equipment and personnel.
Permits of entry (also known as Right-of-Entry Permits) are required for any site exploration outside
of the highway right-of-way whether the site is on private property or on public lands outside the
jurisdiction of ODOT. For simple cases, these permits can be obtained by the geotechnical designer
in charge of the exploration or other staff. For most circumstances however; these permits should be
obtained by the Regions Right-of-Way section. In either case, the region Right-of-Way section should
be consulted prior to any entry onto private property. A sample Permit of Entry Form is included in
Appendix 3-A.
Each permit of entry form should be accompanied by a site map showing the precise location of the
exploration with respect to property lines and any structures or features on the private property.
Considerable delay in the exploration timeline can stem from the permit of entry process. In many
cases, property owners are unaware of upcoming transportation projects until a geologist or
geotechnical engineer asks them for a permit-of-entry for exploration. Even if unopposed or
unaffected by the project, the owner may be reluctant to sign a permit of entry for a variety of
reasons.
Often, further explanation of the activity and its purpose will be all that is necessary, or just allowing
extra time for consideration is all that is required, but will affect the exploration schedule nevertheless.

How to Handle Problems Obtaining Access to Property for Field Investigation


In some cases, landowners are particularly slow in granting access to their property for whatever
reason and may even respond to a request for a permit of entry with a letter from their legal council.
In these instances, the Region right-of-way office should be contacted immediately to take a
lead role in negotiations to resolve the issue. Although the Agency has the statutory authority to
access any real property for the purpose of survey or exploration, it is an exceedingly rare case for

Volume 1
Manual

ODOT Geotechnical Design


9

April 2010

ODOT to exercise this authority for subsurface investigation. The cause for performing a subsurface
investigation on such a property must be well-founded and without feasible alternatives.
Note:
When a property owner refuses permission to enter their property, then all further communication
and resolution becomes the responsibility of the Right-of-Way Section and the project management.
Under no circumstances should field personnel mention or discuss the States statutory authority to
enter upon their property to complete the work, nor should they engage in any bargaining or make
agreements other than those stated on the permit of entry form in exchange for access to their
properties.

Obtaining Right-of-Way from other Real Property-owing Entities


Other real property-owning entities will take more time in granting a permit of entry. Corporations,
governmental agencies, mutually-owned properties, and railroads all have different procedures and
requirements for granting access. Corporations may sign permits of entry only from their main
offices, governmental agencies may have lengthy policies and procedures for granting permissions,
and mutually-owned properties may have numerous non-resident owners that must all be contacted
for their consent.

Railway Right-of-Way
Getting permission to access railroad right-of-way is a special case and can be a particularly timeconsuming undertaking. For local operators and short lines, getting access may be relatively
straightforward. Some larger carriers have a lengthy and rather Byzantine process for handling
permit of entry requests that can severely affect a project timeline. If exploration or access is needed
on railroad right-of-way, the project timeline should be adjusted accordingly and alternatives sought
wherever possible. Permit of entry requests for railroad right-of-way should be forwarded through the
headquarters Right-of-Way section.
In the event that the state-owned railroad right-of-way must be accessed, contact ODOT's Rail
Section to obtain that permit.

Limiting Site Impact


When performing subsurface investigation on private property, all care must be taken to avoid and
mitigate the site impact. Access to such sites should be planned with the smallest possible impact.
Although some exploration sites will be completely removed during construction, there may be
considerable time between then and the time of exploration. The responsibility for complete
restoration of exploration sites is placed on ODOT by the same statute that provides legal access to
those sites.

3 . 3 . 1 . 8 U t i l i t y L o c a t i o n / N o t i f i ca t i o n
Underground and overhead utilities in the project area must be identified and approximately located
early in the Exploration Plan development. The presence of utilities may dictate the location of, or
access to exploration points.
Warning:
Encountering underground utilities during site investigations can be detrimental to the exploration
schedule and budget. Digging or drilling into underground utilities or contacting overhead power

Volume 1
Manual

ODOT Geotechnical Design


10

April 2010

lines with drill rig masts or backhoe arms can be lethal. For these reasons, the exact location of all
utilities must be determined before any equipment is mobilized to the project site.

Utility Notification Center


In Oregon, the law requires that the Utility Notification Center is contacted no less than 48
business hours prior to any ground disturbing operations. This includes all test pit excavation, drilling,
and even hand-augering or digging.
Note:
The Utility Notification Center (or One-Call Center) can be reached at 1-800-332-2344.
The Utility Notification Center contacts all of the utility services with facilities in the location(s)
provided to them based on their records. The individual utilities then dispatch their personnel or
contractors to the site to locate and mark the positions of their facilities according to the instructions
provided. The following occurs in relation to utility marking:

The utilities are also required by law to locate their facilities within 48 business hours. If the
utility operator does not have facilities near the proposed location site, he or she will mark it
as such to indicate that it is safe to proceed. Otherwise, they will mark the approximate
location of their facility in the requested vicinity.

If the utility is close to the proposed exploration, prudence would dictate that the exploration
be moved slightly to allow for errors in the utility location, and to further prevent the accidental
contact with the utility.

If the utility has not marked the requested area in the required time frame, they should be
contacted prior to commencement of exploration to confirm that the utilities have been
contacted, and that they do not have facilities in that area.

The utility operators are often hard-pressed to comply with the 48-hour requirement due to the sheer
volume of utility locations particularly during the summer months when numerous contractors are
requesting them. Additional time may be required, so utility location with respect to projected
exploration starting times should be planned accordingly. It is also important to look for any other
utilities that might be operating in the area in case they are not in the records of the Utility Notification
Center. Indications of other utilities are marked riser boxes, manholes, valves, and obvious
illuminated structures such as street lighting and advertising. It is the responsibility of the project
geologist to notify any other utilities operating in the project area.

Procedures to Perform Prior to calling the One-Call Center


The procedures for utility notification and location are relatively simple, but minor mistakes or
overlooked information can result in unnecessary delay and risk to the utilities and the exploration
personnel. The following steps should be completed and information gathered prior to calling the
One-Call Center:
All proposed exploration sites must be located and clearly marked in the field with a survey lath,
painted target on the ground surface, or both. By convention, the survey lath and target should be
painted white. Efforts should also be made to make the location as visible as possible for the utility
locators such as using additional directional markers and survey flagging.

Each exploration site should be numbered and labeled as either proposed test boring or
proposed test pit.

The nearest physical address or milepost, and the closest cross-street should be recorded.

Volume 1
Manual

ODOT Geotechnical Design


11

April 2010

The Township, Range, and quarter Section should also be determined.

When contacting the One-Call Center, the following information will be asked by their operator:

The callers identification number (one will be assigned if not already registered)

For whom the work is being performed

Who will be doing the work

Type of work

Alternate contact

Location of site (number of exploration points, county, nearest city, address, cross street,
township range, section)

Marking instructions (typically a 25 to 50 radius from each stake or target)

Presence of any overhead utilities

The operator determines which utilities are known to have facilities in that area and provide the list
verbally along with the ticket number which will be used to identify that particular work order. The
operator provides the date and time at which the work should be able to proceed. Once this call is
complete, the operator will then notify those utilities that will then dispatch their locators. ODOT
geotechnical designers use Utility Notification Worksheet, Appendix 3-B, to document utility
location for future reference while on site.

3 . 3 . 1 . 9 M e t h o d s f o r S i t e Ac c e s s
Exploration equipment selected for the subsurface investigation should be matched to the site
conditions. Truck-mounted drills are the most commonly available and are capable of accessing
most sites with or without additional work and equipment. However, for many sites, access to boring
locations can be difficult and even very complex in some cases. Often, the cost for mobilizing special
equipment to a project site is more than compensated for in reduced site impact, reclamation effort,
time and materials costs, and the additional personnel and equipment that might be needed.
Frequently, the method of site access is selected based on one or a combination of desired
outcomes whether time and cost, minimizing impact, equipment availability, or equipment capability.
Truck-Mounted Drill Rigs
Truck-mounted drills that are road-legal generally have limited off-road capability even when
equipped with 4-wheel or all-wheel drive due to their size and weight. These types of equipment are
best suited to work on paved or surfaced areas although they are capable of reaching many off-road
locations in the dry. Because of their axle loading, they can rapidly become mired in wet or soft
soils.
In order to use a truck-mounted drill in difficult conditions, access roads may need to be built using
one or more additional pieces of equipment. In steep terrain, access roads may require substantial
cuts and fills, and where soft ground is encountered, sizeable amounts of rock and geotextile will be
needed to surface the road. Special mats or even plywood may be used to distribute the trucks
weight over soft ground when accessing a boring location. In any case, such work can be expensive,
time-consuming, laborious, and high-impact requiring significant reclamation work after exploration.
Truck-Mounted drills that are off-road capable may require lower-standard access roads, but still
need these roads. If a significant amount of winching or vehicle towing is necessary, an alternative
Volume 1
Manual

ODOT Geotechnical Design


12

April 2010

method of site access should be strongly considered, if only for safety reasons. The advantage of
truck-mounted drill rigs is that they are usually the best-equipped and highest-powered pieces of
equipment available, so if a particular type of drilling or deep hole is required, these may be the only
option. For accessible sites, truck-mounted drills are usually the cheapest and fastest way to
accomplish explorations since they can drive over a site, set up, complete the boring, and move on to
the next location with relative ease and with fewer support vehicles.

Track or ATV-Mounted Drill Rigs


Many exploration drill manufacturers product lines now include drill rigs mounted on a variety of track
and rubber-tire ATV platforms with some of the same features and capabilities as their truck-mounted
counterparts. In some cases, the drilling equipment is the same, and only the platform varies:

Track-mounted drill rigs: Track-mounted drill rigs offer a much greater off-road capability
and ability to access sites in rough terrain and soft ground. Although the track-mounted drill
can reach difficult locations, some road-building or at least clearing of trees and vegetation
may be required, although to a much lesser degree, than their truck-mounted counterparts. A
level pad upon which to set the drill may also need to be constructed. One of the drawbacks
of track-mounted drills is that they require slightly more time for set up and moving between
longer distances since they must be hauled to project sites on a flatbed truck or trailer. The
presence of the trailer or large truck for hauling the drill may also prove to be another
encumbrance when working in tight locations or those sites with limited parking or space for
maneuvering a long truck and trailer combination. The types of tracks must also be
appropriate for the site.
Note:
Older-style steel caterpillar tracks are ideal for traversing steep slopes with a soil cover, but
will be harmful to pavements or landscaped areas. Newer developments with rubber tracks
offer better traction on bare rock surfaces, and are less harmful to pavements and
landscaping but should still be used with caution as their treads can still damage or scar most
surfaces.

ATV Mounts: Typical ATV-mounts consist of balloon or other oversized rubber tires for use
in soft ground or swampy areas. The advantage that such vehicles have over tracks is the
lighter load per unit area and correspondingly reduced impact to sensitive areas such as
wetlands, landscaping, private properties, etc. Because of their distributed load, these
vehicles are more suited to soft or uneven ground applications rather than for sites where
traction on steep slopes is most needed. Several manufacturers now produce ATV platforms
with tractor-style tires that offer many of the advantages of tracked and balloon tires with
respect to traction, impact, and load distribution.

Difficult Site Access


A variety of site conditions and subsurface information requirements create substantial difficulties in
reaching exploration sites whether in remote, environmentally sensitive areas, or restricted space in
the built-up environment. Such obstacles can range from high-angle slopes and physical barriers to
restricted work areas such as confined spaces (as defined by OSHA), limited work space due to
objects or environmentally sensitive areas, and over-water work. Diverse methods are available to
assist with difficult site access as well as drilling contractors that specialize in this type of work.

Volume 1
Manual

ODOT Geotechnical Design


13

April 2010

Methods and equipment for difficult site access are as varied as the sites themselves. The common
factor that limits what methods can be used for certain applications is the weight of the equipment
with the volume of the machinery also being a limitation.

Winching or dragging: Much of this work in the past has been performed by skid or trailermounted equipment with some man-portable also employed in some areas. This equipment
has been winched, crane-lifted, or dragged into place by other tractors.
With the advent of track and ATV-mounted drills, winching and skidding drilling
equipment into place is no longer necessary or recommended due to the amount of
ground disturbance involved.

Cranes: Cranes are often employed to lift equipment into tight work areas although the
weight of many of these drill rigs necessitated very large pieces of equipment to move them
and had their own space issues.

Specialized equipment: Until recently, most of the skid or trailer-mounted and man-portable
drill rigs had restricted power and capabilities. However, drilling technology has advanced to
the point where smaller and lighter equipment is capable of performing heavier drilling tasks.
Specialized difficult-access drilling contractors generally use their own customized equipment
that comes with a specific platform, or breaks down into lighter compartmentalized sections
that are reassembled at the boring location. Much of this specialized equipment is light
enough to be transported while slung beneath a helicopter.

Most modern drilling equipment not mounted on a truck chassis, with the exception of some
man-portable equipment, is capable of completing almost all geotechnical exploration tasks in the
same amount of time as their road-legal counterparts. However, these drills will always be
restricted by allowable axle loads during transport, and so they will always have a disadvantage
with respect to their overall horsepower versus a truck-mounted rig that does not require a truck
and trailer combination for roadway transport. This disadvantage is typically only manifest in very
deep and/or large-diameter boreholes.

Barge/Over-Water Drilling
Foundation investigation for bridges commonly requires in-stream access to drill sites. To achieve
this, barges or other platforms must be used to set the equipment over the foundation location. Overwater work will add extra details to a site investigation, and depending on the location, this can add
extensive logistical complexity to a project.

Permitting: Additional permits will be needed to conduct the over-water work from the
US Army Corp of Engineers and/or the U.S. Coast Guard, and from the port authority or
harbor master with jurisdiction over the waters in which the investigation is being conducted.
An additional staging and launch areas must be identified where equipment can be loaded
onto the barge, and where the crew can access the work site for daily operations. The
appropriate equipment must also be selected for the site with respect to the currents, depths,
river traffic, obstructions, and other details.

Launch site: The site for initially loading and launching the drill barge must be of sufficient
size for the type of equipment being used. The launching ramp should have enough grade to
provide enough draft for the barge. The facility will also need enough room to either drive or
lift the drilling equipment onto the barge and to safely load and unload all other ancillary
equipment and supplies. Scheduling the facility for loading and unloading may also be
important at different times of the year. Some ports may only be available at certain times

Volume 1
Manual

ODOT Geotechnical Design


14

April 2010

due to their ongoing cargo loading operations and public or commercial fishing ramps may be
crowded during those seasons. A more proximate and smaller location may be available for
launching a skiff or other small craft to support the daily drilling operations and permit crew
changes between shifts.

Drilling barge: The barge and any other vessels used for the over-water drilling operations
must also be selected and rigged for the conditions.
o
The drilling barge itself must be of sufficient size not only to support the
weight of the drill and other equipment, but must also have enough deck space
for whatever sampling and testing operations that will also be carried out.
o
The vessel used to transport the drilling barge should also be capable of
moving the barge in all conditions of weather and current.
o
For work in very slow currents or standing bodies of water, the drill barge
may be fixed in place by spud anchors or by lashing to a fixed object such as a
driven pile or pier. Where stronger currents occur, whether stream or tidal, a
larger vessel may be required to transport and anchor the drill barge during
operations. Additional anchoring will be needed in such conditions.
o
Where water levels will fluctuate quickly during the conduct of drilling
such as in tidal zones and downstream of large dams subject to rapid discharge,
allowances must be made for the drill barge to move accordingly with respect to
elevation. These operations will usually require the drill barge to use free-moving
spud anchors that are also fixed to a more securely anchored vessel.
o
The access vessel or skiff must also be capable of operations in all
conditions at the site.
o
Provision must be made for keeping track of elevation changes during
tidal or current changes as this will profoundly affect the drilling operations.

Note:
As a condition of the Corps of Engineers and/or the Coast Guard permit, a licensed Marine
Surveyor must be engaged to examine the equipment and the site conditions. This professional will
then make recommendations concerning the equipment, personnel, and safe conduct of operations.
Whether or not a Marine Surveyor is required, their inclusion for over-water work planning is highly
recommended for the particular skills and efficiencies that they bring to this rather hazardous aspect
of subsurface investigation.

Exploration
Oversight

Management

and

The daily field exploration activities on a project should be based primarily on the execution of the
Exploration Plan. The Exploration Plan provides a framework for scheduling and adjusting field
operations as needed. It will necessarily allow for enough flexibility to modify the subsurface
investigation program as information comes in from the field.

The Project Geologist should maintain a base-level subsurface model from the subsurface
information as it is received in order to make the needed modifications.

Volume 1
Manual

ODOT Geotechnical Design


15

April 2010

The Field Geologist/Drill Inspector will need to provide regular updates on the field activities
and information gathered so that changes to the schedule and routine can be made
expediently. With the advent of cellular telephones and increasing areas of coverage, field
crews should only be a few minutes away from contact with the senior geotechnical
designers to inform them of unanticipated field conditions and in turn, receive direction on
how to proceed with the modifications.

Because of the costs of subsurface exploration and the rapid use of the data, it is imperative that the
subsurface investigation is directly supervised by qualified and experienced personnel. All on-site
personnel including drillers, field geologists/engineers, and testing specialists should be instructed
and familiarized with the project objectives and their role in achieving those objectives. Special
geotechnical or other problems that may be anticipated during exploration including contingencies for
addressing them should also be conveyed. All field personnel should be instructed in their role
concerning project requirements for schedules, environmental protection, and especially, site safety
and health procedures. Field personnel should communicate frequently with project supervisors or
geotechnical designers.
Regular transmission of field data such as boring logs, test data, field conditions, and daily drillers
reports will streamline and economize the site exploration.
Note:
Any unforeseen site changes, complications, and geologic or geotechnical problems revealed during
the investigation that will affect the project scope, schedule or budget should be communicated to the
Project Leader without delay. The geotechnical designer charged with the exploration program is
responsible for immediately and succinctly informing the Project Leader of the nature of the problem,
the expected remediation, and the anticipated impact to the project. The geotechnical designer
should then be prepared to offer alternatives and their respective outcomes for the resolution of the
problem.

3.5.1

Subsurface Exploration Requirements


General

The 1988 AASHTO Manual on Subsurface Investigations is the basis for subsurface investigations
conducted by ODOT. This manual provides guidance on the minimum amount of investigation for
the various structures and geotechnical features constructed for transportation projects. The manual
states however, in numerous places, that there can never be a set of specifications and guidelines
that will determine the amount of exploration that must take place for every project.
Note:
The number of borings, their distribution, sampling interval, and depths of penetration will always be
determined by the underlying geology and the size and complexity of the project.
Planning for the subsurface exploration will be based on past knowledge of the site and on the
published and unpublished literature that was consulted during the project reconnaissance phase.
However, even the most thoroughly studied sites will still reveal previously unknown conditions, and
each exploration provides new information about it. In a sense, the site conditions are truly unknown
until the exploration begins, and knowledge of it increases as the investigation proceeds so
adjustments must be made in the field to economize the investigation while assuring a full
investigation of the important geotechnical design elements.

Volume 1
Manual

ODOT Geotechnical Design


16

April 2010

3.5.2

Exploration Spacing and Layout

The layout of explorations on a project is determined by many variables. As previously discussed,


the assumed complexity of the underlying geology and the type of facility typically dictate the
exploration spacing. Consider the following:

Where conditions are uniform and a considerable amount of previous, reliable work has been
accomplished in a project area, exploration spacing may be increased.

If the geologic conditions are complex and change significantly over short distances, then
explorations will necessarily be conducted on a shorter interval.

Facilities that will impart a heavy load or are more sensitive to settlement or other movements
will also require a more detailed exploration.

The 1988 AASHTO Manual on Subsurface Investigations provides a range of exploration spacing for
the various structures and features that are typically the subject of subsurface exploration.
These guidelines are modified for use within the State of Oregon where subsurface conditions at the
vast majority of sites warrant much tighter exploration spacing due to the highly changeable nature of
the states geology.

3 . 5 . 2 . 1 S p a c i n g a n d L a yo u t S t r a t eg i e s
Because transportation projects are typically linear, explorations tend to be channeled into a relatively
straight and narrow corridor, and are often laid out only along the centerline of many features. This
should be avoided as it most often results in poor development of the subsurface model. To avoid
this, boreholes should be spread out to either side of the centerline to help determine the strike and
dip of the underlying strata, the nature of the contacts (i.e. conformal or non-conformal), and other
changes or irregularities across the subsurface profile. Exploration to reveal or characterize geologic
hazards such as faults and landslides that affect the proposed project may necessarily be conducted
outside of the proposed alignment(s). Material source or disposal site investigations normally take
place far away from the project alignment and will have different exploration spacing criteria.
Take special care when conducting explorations in particular alignments and foundation locations.
Certain geologic conditions, such as openwork cobbles and boulders, heaving sands, or highly
fractured rock may bind exploration tools severely enough that the drill crew is unable to retrieve
them from the hole where they subsequently form an obstruction during drilled shafts construction. In
areas that experience high artesian pressures, improperly sealed boreholes may form an undesirable
conduit for groundwater to enter footing excavations, cut slopes, or cofferdams.
Note:
All borings should be abandoned in accordance to Oregon Water Resources Department
Regulations to prevent vertical water migration. Provision should also be made to extract bound
drilling tools from the boring with special equipment.
The boring layout guidelines presented here are of a general nature and are intended for use in the
preliminary location of site exploration points. The final exploration locations should be developed as
the site investigation proceeds. Information must be incorporated into the Exploration Plan as it
becomes available to assure the most complete, cost-effective outcome.

Volume 1
Manual

ODOT Geotechnical Design


17

April 2010

3.5.2.2 Embankment and Cut Slope Explorations


The maximum exploration spacing for embankment fills over 10 feet (3.05m) in height is 200 feet
(61m). Where changeable conditions or problem areas such as those with soft and/or compressible
materials are present, then the exploration spacing should be decreased to 100 feet (30m). In many
cases it will be necessary to conduct additional exploration using cone penetrometers, hand augers,
or backhoe test pits to further define the properties and boundaries of problem foundation conditions.
At least one boring should be located at the point of maximum fill height.
For cut slopes 10 feet (3m) and higher, the maximum boring spacing is 100 feet (30m). Borings
should be staggered to each side of the cut line to help determine the strike and dip of the units in the
cut slope, and one of the borings should be placed at the maximum depth of the cut. For throughcuts where a cut slope will be located on each side of the roadway, boring spacing may be increased
to 200 feet (61m) for each cut slope, but the borings must be staggered so that the total 100 foot (30)
spacing continues along the length of the cut.
Additional borings will be required in areas of faulted, sheared, tightly folded, highly weathered, or
other potentially detrimental conditions exist.
Hand augers, direct push (i.e., GeoProbe), air-track drills, test pits, geophysical surveys and other
alternative exploration techniques can be used to supplement the test borings in proposed cut slopes
to determine the elevations of variable bedrock surfaces and depths to bedrock. Air-track drills may
also be used to penetrate the bedrock surface to determine and further resolve the location(s) of
weathered rock zones and other features within the proposed cut slope.

3 . 5 . 2 . 3 S u b g r a d e Bo r i n g s
Where relatively unvarying subsurface conditions are predicted and no other foundations or
earthworks are expected, the maximum subgrade boring spacing should be 200 feet (61m). In areas
where highly variably geology is predicted, the boring spacing should be decreased to 100 feet (30m)
and further decreased to 50 feet (15m) in highly erratic conditions. Where critical subgrade
conditions exist, the boring spacing may be decreased to 25 feet (8m).
Alternate exploration methods may be used in variable geologic conditions to supplement the borings
and further resolve the characteristics and distribution of problematic materials and conditions. Such
methods may include hand augers, push-probes, geoprobes, and test pits.

Test pits
Test pits on short intervals (25 feet/8meters) are not recommended due to the potential introduction
of soft areas in the subgrade where the pits were located. If necessary, this problem may be
alleviated by the use of compacted granular backfill materials to abandon the test pits after
exploration. The test pit spoils would then need to be disposed of off-site. Several geophysical
survey methods may also be appropriate for subgrade investigations to supplement the test boring
information. Seismic reflection and electro-magnetic methods are commonly the best-suited for
determining material property boundaries and saturated or water-bearing zones.

3 . 5 . 2 . 4 Tun n e l a n d Tre n ch l e s s P i p e I n s t a l l a t i o n Bo r i n g s
Tunnel construction for highway projects in Oregon is rare; however, trenchless pipe installation is
common. Tunnels and trenchless pipe installations share many common construction and design
issues and are thus treated in a similar manner with respect to subsurface characterization and
exploration. Borehole spacing requirements for tunneling and trenchless pipe installation are highly
Volume 1
Manual

ODOT Geotechnical Design


18

April 2010

dependent on the site geologic conditions and topography. The soil, rock, or mixed-face conditions
predicted will determine the borehole spacing as well as the type of exploration and testing
conducted. The depth of the tunnel/trenchless pipe alignment will greatly influence the total amount
of drilling required.
The actual borehole spacing selected for tunnel or trenchless pipe installation should be determined
by the actual site conditions. These conditions should be identified in advance by preliminary site
review, and in the case of larger projects, preliminary site investigations conducted during the Phase I
field survey. The recommended general borehole spacing for selected conditions is shown in the
following table:

Volume 1
Manual

ODOT Geotechnical Design


19

April 2010

Table 3-10. Tunneling and Trenchless Pipe Installation Recommendations


Recommendations
Soft Ground Tunneling
Adverse Conditions

50-100 feet (15-30m)

Favorable Conditions

200-300 feet (61-91m)

Mixed-Face Tunneling
Adverse Conditions

25-50 feet (8-15m)

Favorable Conditions

50-75 feet (15-23m)

Hard Rock Tunneling


Adverse Conditions

50-100 feet (15-30m)

Favorable Conditions

200-500 feet (61-152m)

Trenchless Pipe Installation


Adverse Conditions

15-30 feet (5-9m)

Favorable Conditions

30-50 feet (9-15m)

In addition to the geologic conditions, other site constraints will equally determine the number and
spacing of borings for tunnels and trenchless pipe installations. The location of existing structures
with respect to the proposed depths and alignments will necessitate a more detailed investigation at
those locations.
Geophysical surveys may also be used in conjunction with the borings to further define the geologic
conditions and to help determine the final boring layouts as defined below.

Wherever possible, horizontal borings should be taken along the proposed tunnel alignment.
Current technology and contractor capabilities allow longer and more accurate horizontal
borings that provide essential information regarding the expected tunnel face conditions.

Trenchless pipe installations through existing embankments can and should be fully
penetrated by horizontal borings to determine the conditions along the full length of the
trenchless installation. Because the horizontal borings do not reveal the conditions above
and below the tunnel/trenchless pipe installation horizons, vertical borings are still required.

Clearly, tunnels with horizontal and vertical curves will be difficult to investigate with horizontal
borings, but as technology advances, methods may soon be available to steer borings along these
alignments.

3 . 5 . 2 . 5 S t r u c t u r e - S p e c i f i c Bo r i n g s
The actual number and spacing for borings for specific structures varies greatly depending on the
predicted geologic conditions and the complexity of the site. In this regard, nearby features such as
streams and environmentally sensitive areas, geologic hazards, and nearby structures will further
prescribe the actual amount of exploration required.
Volume 1
Manual

ODOT Geotechnical Design


20

April 2010

Bridges
For all bridges on ODOT projects, at least one boring will be placed at each bent location. Borings
should be placed at opposite sides of adjacent bent locations when practical as defined below.

For bridges that are 100 feet (30m) wide and larger, at least two borings will be placed at
each bent.

When spread footings are proposed, two borings at opposing corners of the footing are
advisable. Spread footings located on the banks of rivers and streams should be
investigated with at least two borings one on the down-slope and one on the upslope side
of the proposed footing.

If wingwalls greater than 20 feet long are to be constructed, then a boring should be placed at
the end of each wingwall and at 50 foot (15m) intervals from the end of the wingwall to the
bridge abutment.

Trestle-type bridges (usually for detours) should also be investigated at every bent.
Preferably, the borings should be staggered from opposite ends of adjacent bents.

Where highly variable conditions are anticipated, then a boring should be advanced at both
ends of each bent.

For drilled shaft foundations, 1 boring should be placed at the location of each proposed shaft
of 6 feet (1.8m) in diameter and larger. Federal Highway Publication FHWA-IF-99-025
should be consulted for exploration spacing at drilled shaft foundation locations using smaller
diameter shafts.

Culverts
All proposed new and replacement culverts require some level of subsurface investigation as defined
below:

Typically, culverts with a diameter of 6 feet (1.8m) and larger are investigated with test
borings while smaller culverts are investigated with hand-dug test pits or hand auger holes.
However, judgments should be made regarding the actual site conditions and the facility in
question to determine the number and spacing of borings.

Complex geologic conditions merit a more intense investigation, while larger embankments,
adjacent facilities, and proximate unstable slopes may result in a more detailed investigation
for smaller-diameter culverts.

At least two borings should be completed for each culvert up to 100 feet (30m) long.

For culverts longer than 100 feet (30m), borings should have a maximum spacing of 50 feet
(15m).

In complex geologic conditions, boring spacing may be decreased to 20 feet (6m). Borings
will typically be located along the axis of the proposed culvert.

For culvert replacements, the borings should be located immediately outside or partially
within the excavation limits of the original culvert installation with particular care to not locate
a boring where it will penetrate the existing pipe.
Borings will typically be located along the axis of any proposed culvert location.

Volume 1
Manual

ODOT Geotechnical Design


21

April 2010

Box culverts 100 feet (30m) and longer require two borings at each end and at the prescribed interval
between the ends. Refer to Section 3.5.3.4 Tunnel and Trenchless Pipe Installation Borings for
exploration spacing on culverts installed using trenchless technology.

Retaining Walls
Retaining walls higher than 4 feet (1.2m) and any wall with a foreslope and/or backslope angle
steeper than horizontal require a subsurface investigation. At least two borings are required for every
retaining wall regardless of length with the exception of retaining walls less than 25 feet (8m) long.
The maximum borehole spacing along any retaining wall is 100 feet (30m). The preponderance of
retaining walls for ODOT projects will require closer spacing due to the typically variable conditions
encountered. One boring is required at each end of the proposed wall. Where the proposed wall is
longer than 100 feet (30m) long, and less than 200 feet (61m), the third boring may be placed at
either the midpoint of the wall, or at the location of the maximum wall height. Embankments
supported by retaining walls on each side should be investigated as two separate walls.
Borings are typically located on the wall alignment at the proposed location of the wall face however;
they may be staggered to either side of the wall line but should remain within the wall footprint to
evaluate the wall foundation conditions. Consider the following:

For soil nail, tieback, and similarly reinforced walls, additional borings should be completed in
the wall reinforcement zones.

Borings should be located behind the wall in the predicted bond/anchorage zones for tieback
walls, or horizontally 1 to 1.5 times the wall height back from the wall face.

Borings for tiebacks/anchors should be interspersed with the borings along the wall face.
Thus, a 200 foot (61m)-long wall would have (at a minimum) 5 borings 3 along the wall
centerline at the ends and the midpoint and 2 in the prescribed locations behind the wall at
the 50 foot (15m) and 150 foot (46m) points along the wall centerline.

The preceding recommended borehole spacing should be halved for walls that will be constructed to
retain landslides. Landslide retaining walls should have a minimum of 2 borings along the wall line
regardless of length. The maximum borehole spacing along such walls is 50 feet (15m) with
corresponding holes interspersed between located in the bond/anchorage zone. These boreholes
are specifically for characterizing the subsurface conditions at the location of the proposed retaining
wall, and are in addition to any borings advanced to characterize the landslide. Landslide
investigation borings may suffice for the retaining wall investigation only where they fall within the
prescribed locations.

Soundwalls, Traffic Structures and Buildings


Soundwalls and traffic structures, such as mast arm signal poles, strain poles, monotube cantilever
sign supports, sign and VMS truss bridges, luminaire poles, high mast luminaire poles, and camera
poles are common features on highway transportation projects. Buildings such as maintenance
facilities, rest areas, pump stations, water tanks and other unique structures are also sometimes
required for ODOT projects.

Volume 1
Manual

ODOT Geotechnical Design


22

April 2010

Standard drawings have been developed for soundwalls and most of the traffic structures and these
standard drawings contain standard foundation designs for each of these structures. Every
foundation design shown on a standard drawing is based on a certain set of foundation soil
properties, groundwater conditions and other factors that are described on the drawings. These soil
properties and conditions must be met in order to use the foundation design shown on the standard
drawing.
Note:
The subsurface investigation for these structures (with standard foundation designs) should be
sufficient to determine whether or not the subsurface and site conditions meet the requirements
shown on the standard drawings. If the foundation conditions at the site are determined not to meet
the subsurface and site conditions described on the standard drawings (e.g., poor soil conditions or
steep slope), then the standard drawings cannot be used and a site specific foundation investigation
and design is required.
For buildings and traffic structures without standard foundation designs, the foundation conditions
must be investigated sufficiently to determine the soil properties and groundwater conditions required
for a site specific foundation design.
All new soundwalls, traffic structures or buildings require some level of subsurface investigation.
Considerable judgment is needed to determine which structures will need site-specific field
investigations and the extent of those investigations. If the available geotechnical data and
information gathered from the site reconnaissance and/or office review is not adequate to make an
accurate determination of subsurface conditions, then site specific subsurface data should be
obtained through a proper investigation. In these cases, explorations consisting of geotechnical
borings, test pits and hand auger holes, or a combination, shall be performed to meet the
investigation requirements Most of these structures require site-specific soil and/or rock properties for
their foundation design or to determine foundation embedment depths when using a standard
drawing. Therefore, subsurface exploration should be anticipated for most soundwall, traffic
structure, and building foundation designs.
The extent of the investigation will be largely dependent on the predicted site conditions and the type
of structure. At unfavorable locations, drilling and sampling may need to be conducted more
frequently while sites with favorable conditions may allow for less frequent and/or less expensive
investigation methods such as hand augers holes and test pits.
As a minimum, develop the subsurface exploration and laboratory test program to obtain information
to analyze foundation bearing capacity, lateral capacity, stability and settlement.
The following information is generally obtained:

Geological formation(s)

Location and thickness of soil and rock units

Engineering properties of soil and rock units such as unit weight, shear strength and
compressibility

Groundwater conditions (seasonal variations and maximum level over the design life of the
structure)

Ground surface topography

Local considerations, (e.g., slope instability potential, expansive or dispersive soil deposits,
utilities or underground voids from solution weathering or mining activity)

Volume 1
Manual

ODOT Geotechnical Design


23

April 2010

Specific field investigation requirements for soundwalls, traffic structures and buildings are
summarized in Table 3-11. Specific field investigation requirements Note that the term borings in the
table refers to conventional geotechnical boreholes while the term exploration points may consist of
any combination of borings, test pits, hand augers, probes or other subsurface exploration device as
required to adequately determine foundation conditions.
Table 3-11. Specific field investigation requirements
Structure Type
Mast Arm Signal Poles,
Strain Poles,
Sign and VMS Truss
Bridges, Monotube
Cantilever Sign Supports,
Luminaire Poles, High Mast
Luminiare Supports, and
Camera poles.

Field Investigation Requirements

For mast arm signal pole or strain pole foundations within


approximately 75 ft of each other or less, such as at small to
moderate sized intersections, one geotechnical boring for the
foundation group is adequate if conditions are relatively uniform.
For more widely spaced foundation locations, or for more variable
site conditions, one boring near each foundation should be
obtained.

Investigate sign and VMS truss bridges with one boring at each
footing location unless uniform subsurface conditions are
sufficient to justify only a single boring. Where highly variable
conditions occur or where the sign bridge footing is proposed on
a slope, additional borings or exploration points may be
necessary.

For single, isolated monotube cantilever signs; one geotechnical


boring at each footing location.

Luminaires, High Mast Luminaire Supports and Camera Poles; one


exploration point at each footing location.

The depth of the explorations should be equal to the maximum


expected depth of the foundation plus 2 to 5 ft.

Sound Walls

For sound walls less than 100 ft in length, a geotechnical boring


approximately midpoint along the alignment and should be completed on
the alignment of the wall. For sound walls more than 100 ft in length at
least 2 borings are required. Borings or exploration points should be
spaced every 100 to 400 feet, depending on the uniformity of subsurface
conditions. Where adverse conditions are encountered, the exploration
spacing can be decreased to 50 feet. Locate at least one exploration point
near the most critical location for stability. Exploration points should be
completed as close to the alignment of the wall face as possible. For sound
walls placed on slopes, an additional boring off the wall alignment to
investigate overall stability of the wall-slope combination should be
obtained.

Building
Foundations

The wide variability of these projects often makes the approach to the
investigation of their subsurface conditions a case-by-case endeavor. The
following minimum guidelines for frequency of explorations should be
used. More detailed guidance can be found in the International Building Code
(IBC). Borings should be located to allow the site subsurface stratigraphy
to be adequately defined beneath the structure. Additional explorations
may be required depending on the variability in site conditions, building
geometry and expected loading conditions. Water tanks constructed on
slopes may require at least two borings to develop a geologic crosssection for stability analysis.

Volume 1
Manual

ODOT Geotechnical Design


24

April 2010

Table 3-2 (Cont.)


Structure Type

Field Investigation Requirements


Building surface
area (ft2)

No. of Borings
(minimum)

<200

200 - 1000

1000 - 3,000

>3,000

34

The depth of the borings will vary depending on the expected loads being
applied to the foundation and/or site soil conditions. All borings should be
extended to a depth below the bottom elevation of the building foundation
a minimum of 2.5 times the width of the spread footing foundation or 1.5
times the length of a deep foundation (i.e., piles or shafts). Exploration
depth should be great enough to fully penetrate soft highly compressible
soils (e.g., peat, organic silt, soft fine grained soils) into competent material
suitable for bearing capacity (e.g., stiff to hard cohesive soil, compact
dense cohesionless soil or bedrock).

In addition to the exploration requirements in Table 3-11. Specific field investigation requirements,
groundwater measurements, conducted in accordance with Chapter 3, should be obtained if
groundwater is anticipated within the minimum required depths of the borings as described herein.

3 . 5 . 2 . 6 C r i t i c a l - Ar e a I n ve s t i g a t i o n s
In areas where critical geologic conditions or hazards such as highly irregular bedrock surfaces,
extremely weathered or altered rock, compressible materials, and caverns or abandoned
underground facilities are predicted from detailed background study or preliminary exploration, it may
be necessary to further investigate the area with additional explorations. Such investigations
normally involve drilling on a grid pattern over the area in question. An initial, wider grid pattern may
be selected to locate the area of most concern with a closer grid pattern used later to further
characterize the area of concern. Grid pattern investigations may consist of hand auger holes, direct
push holes, or cone penetrometers in addition to the more conventional test borings. Geophysical
surveys may also be used to establish or refine the boundaries of the grid pattern investigation.

3.5.2.7 Landslides
The number and layout of test borings for landslide investigation depends upon the size and nature
of the landslide itself and on the results of detailed site mapping and initial subsurface models based
on the mapping. Since information about the subsurface is unknown initially, landslide investigation
largely becomes an iterative process as new data obtained provides information that is used to
further develop enough knowledge of the landslide to begin stability analysis.
The approach to landslide investigation is very complex and involves numerous techniques and
procedures, and is discussed in greater detail in Chapter 13. This chapter is intended to convey a
general sense of the layout of the borings needed for a typical landslide investigation.

Volume 1
Manual

ODOT Geotechnical Design


25

April 2010

Enough borings must be made initially to fully develop at least one geologic cross-section through the
axis of the slide. Consider the following:

As a minimum, there should be borings near the top, middle, and bottom of a known or
potential landslide area. Ideally, the borings would be placed in the toe or passive wedge
area (if applicable), at the head or active slide zone, the area of transition between the active
and passive zones, and in the areas behind the headscarp and in front of the toe outside of
the slide zone.

For longer slides, space additional borings in the active and/or passive slide zones on 50 foot
(15m) intervals.

Place additional borings on a 50 foot (15m) interval in a line perpendicular to the direction of
slide movement at the deepest zone of slide movement.

For investigation of areas of potential slide movement, a grid pattern of explorations are usually
selected for preliminary identification and delineation of the affected area. The grid spacing is
dependent on several factors. Usually, the predicted size of the landslide, results of remote
sensing, availability of previous data, and site access will primarily determine the spacing
between borings. Where large areas would potentially be affected by landslide movement, a 200
foot (61m) square or staggered grid spacing is sufficient for preliminary identification.

Subsurface Investigations on Unstable Rock Slopes


Subsurface investigations for unstable rock slopes are necessary when a significant amount of rock
excavation is needed to accommodate highway realignment or an increased fallout area.

Typically, the amount of information available at a large, accessible rock exposure is sufficient
for minor slope modification, and of generally greater value than core drilling with respect to
information concerning rock conditions.

However, when significant modification of the slope is considered for realignment and/or
rockfall mitigation, subsurface investigation is frequently needed to determine the rock
character within the proposed cut, overburden thicknesses, groundwater conditions, threedimensional character of the units (if unknown), and other important design and construction
information.

Drilling is recommended to assure continuous subsurface conditions throughout the


excavated rock material.

The skilled geologists interpretation of the outcrop generally provides enough information for rock
slope design, but the changeable nature of the states geology, and the need to assure subsurface
conditions to prevent construction delays and claims is usually reason enough to gain the additional
assurance of further subsurface data. This is not to state that drilling for a rock cut slope modification
is automatic. The geotechnical designer must determine the cost-benefit of additional subsurface
investigation based on the local geology and the risks involved.

Volume 1
Manual

ODOT Geotechnical Design


26

April 2010

Note:
For the assessment of large block or wedge failures, subsurface investigation should proceed in a
similar manner to the approach to landslide investigations as described above. Some of the borings,
or additional borings may be needed at prescribed orientations other than vertical to assess the
projected failure planes.
For projects where realignment or slope modification to increase the fallout area is needed the
investigation should carry on according to the procedures for cut slope investigation described in
Section 3.5.3.2.

3.5.3

Exploration Depths

Determining the required depths of subsurface explorations requires the consideration of many
variables such as the size, type, and importance of the structure, and most of all, the underlying
geology. Consider the following:

The borings should penetrate any unsuitable or questionable materials and deep enough into
strata of adequate bearing capacity where significant settlement or consolidation from the
increased loads from the proposed structure is reduced to a negligible amount. The stress at
depth added by the structure is usually taken from the appropriate tables and charts or
determined using the Boussingesq or Westergaard solutions.

All soft, unsuitable, or questionable strata should be fully penetrated by the borings even
where they occur below an upper layer of high bearing capacity.

Test borings should not be terminated in low-strength or questionable materials such as soft
silt and clay, organic silt or peat, or any fill materials unless special circumstances arise while
drilling.

3 . 5 . 3 . 1 Ter m i n a t i o n D e p t h s
When competent bedrock is encountered, test borings may generally be terminated after penetrating
15 feet (4.5m) into it. Where very heavy loads are anticipated, test borings may be extended to a
considerable depth into the bedrock depending on its characteristics and verification that it is
underlain by materials of equal or greater strength. For most structures, it is advisable to extend at
least one boring into the underlying bedrock even when the remaining borings are terminated in soils
of adequate bearing capacity.
As with all other aspects of subsurface investigation, considerable professional judgment is needed
to determine the final depths of planned explorations. Generally, previous subsurface information is
needed to determine the approximate depth of the proposed borings on the Exploration Plan. Where
this information is unavailable, general guidelines can be used to establish the preliminary exploration
depths and quantities. These guidelines are outlined for specific geotechnical features in the
following sections.

3 . 5 . 3 . 2 E m b a n k m e n t a n d C u t S l o p e E x p l o r a t i o n D ep t h s
For embankments of 10 feet (3m) or greater in height, the test borings should penetrate from 2 to 4
times the proposed fill height or more depending on the final width of the roadway and the actual
materials encountered. If suitable foundation materials are encountered such as dense granular soils
or bedrock, the depth may be decreased up to a minimum depth equaling the height of the
Volume 1
Manual

ODOT Geotechnical Design


27

April 2010

embankment. Where confined aquifers with artesian pressures or liquefiable soils are present, the
exploration depth should be extended to fully penetrate these units.
Cut slopes with a depth of 10 feet (3m) or more should be explored to a depth that is two times the
height of the proposed cut. When bedrock is encountered in a cut slope boring, the boring should
extend at least 15 feet below the finish grade of the cut. Cut slope borings should be extended if
sheared surfaces or other evidence of landslide susceptibility are encountered that could affect the
performance or constructability of the finished slope.

3 . 5 . 3 . 3 S u b g r a d e Bo r i n g s
Where minor amounts of earthwork (cut slopes less than 10 feet (3m) deep) for the alignment profile
are expected, test borings and test pits should extend 15 feet (4.5m) below the proposed final grade
elevation. Where bedrock or other hard materials are encountered, coring should be extended 15
feet (4.5m) into the hard stratum to evaluate their conditions. For fill areas less than 10 feet (3m)
high, explorations should extend to 15 feet (4.5m) below the original ground surface unless
questionable materials are encountered. If soft, organic or other deleterious materials are
encountered in subgrade borings, the depth of exploration should be increased as necessary to fully
evaluate those materials.

3 . 5 . 3 . 4 Tun n e l a n d Tre n ch l e s s P i p e I n s t a l l a t i o n Bo r i n g s
A rule-of-thumb for tunnel exploration is the amount of exploration drilling should be 1.5 times the
length of the tunnel. This should be considered as a bare minimum for exploration cost estimating for
tunnel/trenchless installation projects will shallow alignments in very favorable conditions, and does
not include horizontal drilling along the tunnel/pipe profile. Clearly, the amount of drilling for any given
length of tunnel/trenchless installation alignment is dependent on several factors that include, among
others, the depth of the invert, diameter of the tunnel/pipe, geologic conditions, and contingencies.
Typically, tunnel/trenchless installation borings should be extended at least 1.5 tunnel/pipe diameters
below the proposed grade of the invert. It may be beneficial to further extend the borings to as much
as 3 times the tunnel/pipe diameter as a contingency if the final tunnel alignment has not been
determined. The depth of the borings should be increased further to evaluate any unforeseen or
unfavorable geologic conditions encountered that may impact the tunnel or pipe design and
construction. Wherever practical, horizontal borings should be taken along the tunnel profile because
of the advantages of having a full-length representation of the actual tunnel/pipe horizon conditions.

3 . 5 . 3 . 5 S t r u c t u r e - S p e c i f i c Bo r i n g s
The guidelines for boring depths presented in Section 3.5.3 stem from structure-specific boring
guidelines developed by AASHTO and other agencies. Follow these guidelines:

It is highly desirable for all structure-specific borings to penetrate at least 15 feet (4.5m) into
bedrock.

For drilled shaft installations, the test borings should be advanced 1.25 times the total
projected shaft length beyond the predicted shaft base elevation.

If the shaft base is to be founded in soil or rock with an RQD of 50% or less, then the test
borings should be extended an additional depth below the proposed bottom of the shaft equal
to the larger of 20 feet (6m) or 3 times the shaft base diameter. Shafts are most commonly
designed to bear on competent bedrock, thus, where the RQD is greater than 50%, the test

Volume 1
Manual

ODOT Geotechnical Design


28

April 2010

boring should also be advanced to the greater of 20 feet (6m) or 3 times the shaft base
diameter below the estimated shaft base elevation.
Note:
The geotechnical designer must exercise judgment concerning the nature of the facility with respect
to the total and economical amount of drilling needed for the specific structure. Borings for
soundwalls, small traffic structures or culverts may not be required to obtain core samples in
bedrock, but for bridge foundations, bedrock drilling would certainly be needed.

3 . 5 . 3 . 6 C r i t i c a l - Ar e a I n ve s t i g a t i o n s
In those areas where unfavorable or critical geologic conditions are expected to have an adverse
effect on the project design and construction, the explorations should be extended to a depth where
those conditions may be fully evaluated. All problematic strata and areas of concern should be fully
penetrated by the borings. It is advisable to extend the borings beyond the depths that are strictly
necessary rather than terminate them before the desired information is obtained. Borings should
never be terminated in soft, organic, or any other deleterious materials that will adversely affect the
project design, construction, or performance. Extra drilling in some borings is less expensive than
drilling additional borings or even remobilizing equipment to the site to obtain sufficient data for
design.

3.5.3.7 Landslides
Considerable flexibility must be built into the Exploration Plan for any landslide, and particularly with
respect to the depth of the explorations. Follow these guidelines:

Typically, the cross-section drawn along the centerline of the landslide is used to develop the
preliminary exploration depths.

Circular, elliptical, or composite curves drawn from the headscarp to the toe bulge are
projected onto the cross-section to show the possible depths of slide movement. These
curves are commonly exaggerated to conservatively estimate the slide depth.

The preliminary boring depths should extend 20 feet (6m) or more below the projected slide
plane to assure that the zone of movement is fully penetrated, and to secure instruments
below the slide plane for the best results.

Firm, resistant strata, bedrock projections and irregular surfaces will also affect the geometry
of the slide plane, and subsequently, the final depths of individual borings.

Landslide borings should always be extended to a depth that clearly identifies which
materials are involved in the current slope movement, which underlying materials are
presently stable, and the location of the slide surface(s). This is not only important to the
development of a stability analysis, but will become important once again during construction
when the precise locations of mitigation efforts will be determined. There is often a possibility
that the observed landslide activity is an accelerated portion of a slower, deeper-moving
landslide that may only be detected by instrumentation. For this reason, at least one boring
should be extended far below the predicted slide surface to divulge such activity. Any
Exploration Plan for landslide investigations should contain the flexibility to extend borings to
considerable depth during the site exploration.

Volume 1
Manual

ODOT Geotechnical Design


29

April 2010

3.5.4

Sampling Requirements

Since the primary purpose of the subsurface exploration program is the collection of samples that are
as closely representative of actual site conditions, the sampling requirements are typically the most
stringent in the Exploration Plan. Particular care must be taken in their method of collection,
measurement, handling, and preservation since field and laboratory testing results are so greatly
dependent on the quality of the sampling. Sampling requirements are also subject to the same
variables that affect exploration layout and depth.

Sampling interval: Most Exploration Plans will have a set maximum sampling interval. For
most ODOT projects, Standard Penetration Tests (SPTs) are taken, and samples retained, on
2.5-foot (0.76m) intervals in the first 20 feet (6m) of the boring, and on 5-foot (1.5m) intervals
thereafter to the bottom of the hole or until rock coring begins. In addition to this minimum
interval, samples should also be taken at each noted change in material or subsurface
condition. Where thick, uniform strata exist, a wider sampling interval may be warranted
however, this greatly depends on the extent of previous site knowledge and project
requirements. Where complex conditions and/or numerous strata exist, the sampling interval
may be increased to a shorter sampling interval.

Sample collection: Samples should be collected from each identified stratum, preferably
from more than one boring to fully characterize each unit. In addition, undisturbed samples
should be obtained from all cohesive soil units encountered. It is frequently warranted to drill
additional borings to obtain undisturbed samples in particular units that may have been
missed by previous sampling intervals or to further characterize those units. Where a larger
volume sample is needed, a variety of sampling methods and techniques can be utilized
including oversized split-spoons, various coring methods, and Becker-hammer drills.
Sampling techniques are discussed in the next section.

Continuous sampling: Continuous sampling is beneficial in areas of changeable site


conditions and underlying geology as well as critical zones for project design. The zones
immediately below proposed foundation elevations should be sampled continuously in
addition to the zones immediately above, through, and below projected landslide zones of
movement. For tunnel/trenchless pipe installations, continuous sampling should be
conducted for 1 tunnel diameter above and below the tunnel horizon as well as the tunnel
horizon itself. Soil and rock coring is by its nature, a continuous sample, and is the most
common method to obtain a continuous representation of the subsurface materials.
However, continuous SPTs, Shelby Tubes, or a combination of these and other methods can
be used.

Observation: Careful observation and evaluation during drilling and logging of the recovered
samples is essential to the entire exploration program. Much information can be recovered
even when sample recovery itself is minimal.

3.5.5

Sampling Methods

Various sampling methods are described in this section. Many of the sampling methods are based on
ASTM International standards located at www.astm.org (the ASTM Site).

Volume 1
Manual

ODOT Geotechnical Design


30

April 2010

3 . 5 . 5 . 1 S t a n d a r d P e n e t r a t i o n Test i n g
All Standard Penetration Tests must be performed according to ASTM D 1586-99. The Standard
Penetration Test (SPT) is the most common method for field testing and sampling of soils. Some
variations with respect to standard intervals and refusal criteria occur throughout the industry
however the fundamental procedure still adheres to the ASTM standard. The SPT uses the following
methods:

This sampling method uses the standard configuration 2-inch (5cm) outside diameter split
spoon sampler at the end of a solid string of drill rods. The split spoon is driven for a 1.5-foot
(0.45m) interval using a 140 Lb (63.5 Kg) hammer dropped through a 30-inch (76cm) free
fall.

The number of hammer blows needed to advance the sampler for each 6-inch (15cm)
interval is recorded on the boring log and sample container.

The Standard Penetration Resistance or uncorrected N-value is the sum of the blows
required for the last two 6-inch (15cm) drives. Refusal is defined as 50 blows in 6 inches
(15cm) of penetration and recorded on the log as 50 blows and the distance driven in that
number of blows.

The hole is advanced and cleaned out between sampling intervals for at least the full depth of
the previous sample.
This general procedure can be used with larger diameter samplers and heavier hammers for the
purpose of obtaining additional sample volumes, but the blow counts do not provide standard
resistance values. Prior to the commencement of drilling operations, the hammer energy must be
measured to determine the actual hammer efficiency. This information can usually be obtained by
the drill manufacturer. If it is not available, a competent technician must be engaged to measure the
hammer energy for each drill rig.

3 . 5 . 5 . 2 T h i n - Wal l e d U n d i s t u r b e d Tub e S a m p l i n g
Undisturbed samples of cohesive soils should be taken with 3-inch (7.6cm) diameter Shelby Tubes
according to the standard practice for thin-walled tube sampling of soils in
ASTM D 1587-00. This method obtains relatively undisturbed samples by pressing the thin-walled
tube into the subject strata at the bottom of the boring. Thin-walled sampling is simply a method for
retrieving a sample for laboratory testing. There is no actual field testing involved with thin-walled
sampling unless a Torvane or Pocket Penetrometer test is performed on the end of the sample.
Pressures exerted by the drill rig while pushing Shelby tubes are frequently recorded for general
reference but do not provide repeatable test results. After the unfavorable effects of the sampling
procedure, transport, handling, and storage, a truly undisturbed sample cannot be realistically tested
in the laboratory. However, with appropriate care, valid samples can be taken for shear strength,
density, consolidation, and permeability testing.
Shelby tubes do not utilize a sample retention system to hold the sample in place during retrieval
from the borehole, so sample recovery can be unreliable. Thin-walled sampling in general is
successful only in soft to stiff cohesive soils. Soils that are very soft are difficult to recover with
standard Shelby tube while the upper range of stiff and very stiff soils are difficult to penetrate or bend
the tube resulting in a disturbed sample. Oversized clasts and organic fragments in the softer soil
matrix can also be detrimental to thin-walled sampling.

Volume 1
Manual

ODOT Geotechnical Design


31

April 2010

Various samplers that use retractable pistons to create a vacuum in the top of the tube can achieve
greater success in obtaining undisturbed samples of soft cohesive soils as well as granular materials.

3.5.5.3 Rock Coring


Rock core drilling should be carried out according to ASTM-D 2113-99. Successful core drilling is as
much a skill as it is a test procedure. Experienced, conscientious personnel are necessary not only
to run the equipment, but also to interpret the results of the drill action as well as the samples
recovered. Material recovered may not actually represent the subsurface conditions present if not
correctly sampled. Observation and interpretation of the drill action, fluid return, and other
characteristics provide indications of the actual validity of the core sample as well as other
information concerning the actual conditions in the subsurface.
Note:
ASTM states that the instructions given in D 2113-99 cannot replace education and experience and
should be used in conjunction with professional judgment. Qualified professional drillers should be
given the flexibility to exercise their judgment on every alternative that can be used within the
appropriate economic and environmental limitations.

Triple-tube Core Barrel Systems


Because of the close-jointed, highly fractured nature of many rock formations in Oregon, and the
detailed observations desired, rock coring should be performed with triple-tube core barrel systems
that are best-suited to such material. These systems provide the best recovery in difficult, highly
fractured and/or weathered rock which is extremely important since discontinuity spacing and
weathering characteristics usually limit the strength of a rock mass with respect to foundation loading,
or the performance of rock excavations. Triple-tube barrels provide direct observation of the rock
core specimen in the split-half of the innermost tube as it is extracted from the inner core barrel. This
allows accurate measurement of RQD and recovery and discontinuity attitudes prior to further
specimen handling. Partial isolation of the sample in the inner split-barrel from the drilling fluids also
preserves much of the discontinuity texture and infilling material that is also very important to rock
mass characterization.
Most rock coring is performed with H-sized systems that provide core specimens with a diameter of
213/32 inches (61.1mm).
Note:
Considerable penalties occur with respect to sample quality when using smaller diameter coring
systems due primarily to drill action, particularly at greater boring depths; thus, H-sized core should
be considered the minimum size for explorations.
Larger diameter cores also provide a better assessment of discontinuity properties. There may be
situations where smaller diameter coring is necessary such as difficult access sites where small
equipment is needed that may not have the torque required to turn larger diameter casing. Core runs
are typically made in 5-foot sections since this is the approximate length of most commonly-available
core barrels. Runs may be shortened when difficult drilling conditions are encountered. Longer
barrels may also be used in highly favorable conditions such as quarry site investigations or other
areas with uncommonly massive rock.

Volume 1
Manual

ODOT Geotechnical Design


32

April 2010

Rock core specimens should be preserved and transported according to the standard practice in
ASTM D 5079-02. Core specimens should always be extruded from the inner core barrel using the
hydraulic piston system. The inner split barrel should not be manually rammed out of the inner barrel
as this will result in sample disturbance. The core should not be dumped out of the end of the barrel
either since this will also disturb the sample as well as invalidate some of the information.

3.5.5.4 Bulk Sampling


Bulk sampling should be carried out at all pipe/culvert locations from the actual invert elevation when
test borings are not required. The samples collected are submitted for the appropriate chemical
testing. Typically, bulk samples of 25 lbs (11Kg) if impermeable bags are used, or 2 gallons (7.5
liters) for jar/bucket samples are collected from each discrete sampling site. Sample receptacles
must be sealed to preserve natural moisture conditions. Bulk sampling may also be conducted for
material source investigations and other surficial applications. All samples collected should be
preserved and transported according to ASTM D 4220-95.

3.5.6

Sample Disposition

Soil and rock samples collected during subsurface exploration should be transported to the
appropriate ODOT region storage facility upon completion of the investigation. Soil samples are
usually retained for only a short period of time after project construction since physical and chemical
changes occur that, over time, invalidate the results of further testing regardless of any effort to
preserve them. Rock core specimens are typically retained for 3 years after the final acceptance of
the project or when the contractors and other concerned parties have been settled with provided that
there are no problems with the performance of the facility. Specimens related to future construction
activities should be retained. Under no circumstance will soil samples and rock core specimens that
may have a bearing on an unsettled claim be disposed of until such claims are finally resolved.

3.5.7

Exploration Survey Requirements

The actual location and elevation of all exploration sites should be surveyed and plotted on the
project base map. Once exploration is complete, the actual exploration site should be marked with a
survey lath or painted target so that the survey crew can readily measure the intended location. The
exploration number should also be marked in the field for accurate reference by the surveyors.
Surveys should be completed based on the project coordinates in addition to the WSG-84 datum.
Elevations should be referenced to Mean Sea Level (MSL).

3.6.1

Subsurface Exploration Methods


General

Many factors influence the applicability and selection of subsurface exploration equipment and
methodology for any selected project site investigation. Selection of equipment and methods are
usually based entirely on geotechnical data needs and geologic conditions but may also be based on
site access, equipment availability, project budget, environmental restrictions, or a combination of any
of these.
In many cases, trade-offs between expected results and the exploration method chosen must be
evaluated to achieve the needed results within defined time limits and project budget constraints.

Volume 1
Manual

ODOT Geotechnical Design


33

April 2010

Geotechnical designers should be familiar with the exploration methods applied on their projects, and
their results and potential limitations or effects on the data they receive from the field.
Most test borings conducted for transportation projects in Oregon are standard diameter vertical
borings using rotary or auger drilling methods. Sampling within the boring is typically done by
Standard Penetration Tests (SPTs), 3-inch (7.62cm) Undisturbed Shelby Tube samples, HQ3-sized
rock coring, and auger coring. Additional, supplementary explorations are conducted using hand
augers, direct push (i.e. GeoProbe) rigs, cone penetrometers, and test pits dug either by hand or
more commonly with hydraulic excavators. ODOT is currently evaluating and using newer
exploration technologies as they are developed or become increasingly available. The use of sonic
drilling and geophysical methods are examples.

3.6.2

Test Boring Methods

The most commonly used drilling methods on ODOT projects are auger boring and rotary drilling.
Continuous sampling core drilling is employed with both methods. Most modern drill rigs are capable
of employing both of these techniques with only minor adjustments to the tooling in the field. Other
techniques that are less commonly used are displacement borings using rotosonic or percussion
methods. Each drilling method should be selected based on the quality of information obtained in the
materials for which the drilling method is best suited for, thus, selection of drilling technique should be
carefully considered. Since most test borings penetrate many types of materials, several techniques
are commonly employed in any single test boring. Various institutions or individuals have strong
preferences for certain types of drilling methods and will tend to use them as a default for almost
any condition encountered. This behavior should be corrected or avoided. Almost every technique is
capable of penetrating the subsurface or making a hole. The quality of the results is the purpose of
subsurface investigation, and different drilling techniques are better suited to certain materials and
conditions. Achieving quality results from a drilling program are more important than convenience.

3 . 6 . 2 . 1 M e t h o d s G e n e r a l l y No t U s ed
Cable-tool, wash, jet, and air-rotary methods are generally not used on ODOT projects for many
reasons. Cable-tool drilling may be useful for some environmental applications and well
installations, but is generally antiquated and not productive for geotechnical investigation. Wash
and jet borings cause down-hole disturbance well past the bottom of the boring, and the fluids
are difficult to recover making them more of a liability than a source of data. Air-rotary drilling
usually causes too much down-hole disturbance to provide reliable SPT data, and difficult to
advance in soft soils. Groundwater typically stops further advancement of air-rotary drills, forms
large voids, and casts sediment-laden water about the site. Air-rotary drilling may be suited to
specific applications where known materials at a site are delineated based on the drill advance
rate and obvious changes in the drill cuttings as they are flushed from the hole. In these
applications, the air-rotary borings should be supplemental to standard geotechnical exploration
borings conducted at the site.

Volume 1
Manual

ODOT Geotechnical Design


34

April 2010

3 . 6 . 2 . 2 Au g e r Bo r i n g s
Rotary auger drilling is one of the more rapid and economical methods of advancing exploration
borings. Most modern drilling equipment has enough power to turn augers of considerable diameter
to a substantial depth. Currently, most augering uses a hollow-stem auger that allows the hole to
remain cased while the various sampling or drilling tools are used and withdrawn from the hole with
drill rods or wireline retrievers. A central stinger bit or plug is placed at the bottom of the auger while
the boring is advanced. Solid stem auger use has largely been discontinued due largely to the
advent of hollow stem augers and the more powerful equipment that is capable of turning their larger
diameter drill string. The standard practice for using hollow-stem augers is described by ASTM D
6151-97. Auger boring has many advantages and disadvantages for various materials encountered
as described below.

Auger Boring Advantages


Auger boring has many advantages and disadvantages for various materials encountered. The
primary advantages of augers are the preservation of the natural moisture content of the soil and the
rapid advancement of the drill through soft to stiff soils. Augers are also useful where drill fluids are
difficult to obtain or are an environmental concern, and in freezing conditions where the use of water
is problematic. An additional advantage of augers is that they create a large enough hole to install
larger-diameter standpipe piezometers or nested piezometers in conformance with Water
Resources Department regulations. In addition, the natural piezometric surface is more readily
monitored during drilling. Coring tools are also available for auger systems that provide continuous
sampling in soils and even weak rock materials. These tools can be placed by either rods or wireline
into special auger bits that feed a continuous soil sample into a split barrel that is then retrieved in 2.5
or 5-foot (0.76-1.52m) sampling intervals. Plastic liners that fit in the auger core barrel can also be
used to preserve soil cores in their natural moisture conditions.

Auger Boring Advantages and Disadvantages


The disadvantages of augering are the power needed to turn long strings of auger in dense
formations, the volume of the hole and the cuttings created, and the disturbance of the natural
materials in certain conditions. When hollow-stem augers are used in granular soils below the water
table, the hydrostatic pressure differential between the inside and outside of the auger casing will
force saturated sands, silts, and fine gravels up into the casing effectively loosening the materials
below the auger bit. This can be caused by either the natural differential, or by the pressure induced
during retraction of the stinger bit or plug. The augers themselves can also affect the conditions of
loose granular materials and silts ahead of the bit. In both cases, SPT values obtained will be
different than what is true for the natural conditions. To counter this effect, a head of water or other
drilling fluid can maintained in the auger casing to counteract these effects. Adding fluids to the auger
generally negates their advantages and if such action is necessary, a different drilling technique
should be employed. Hollow stem augering should not be employed when assessing liquefaction
potential.
A common complaint about augering is the volume of cuttings generated. Where disposal is a
concern, this is probably a disadvantage. However, when drilling in an environmentally sensitive
area, augering is often preferable because the cuttings are easily contained on site when drilling
above the water table. A past complaint has also been the weight of the augers themselves although
this has largely been negated by the more powerful equipment and the available wire line systems to
assist with moving them around the site.

Volume 1
Manual

ODOT Geotechnical Design


35

April 2010

3.6.2.3 Rotary Drilling


Rotary drilling is the most common, and usually the most versatile drilling method available. Various
tools and products available for rotary drilling allow it to be adaptable to most drilling conditions and
geologic materials. Rotary boreholes can be uncased holes advanced with a drill bit on rods or
cased holes made with a casing, casing advancer and casing shoe. The casing advancer is a driver
assembly with latches that fit in the bottom of the casing where it holds the center bit at the bottom of
the hole and is subsequently retrieved with a wireline system. This method of drilling involves a
relatively fast rotation speed, fluid circulation and variable pressure on the drill bit to penetrate the
formation, pulverize the formation particles at the bottom of the borehole. The circulating fluids carry
these cuttings away from the bit, up the borehole annulus, and out of the hole.
When the desired sampling depth is reached, the drill rods or casing advancer are retracted from the
hole and replaced with the desired sampling tool. The sampling/testing is conducted while the hole is
filled with fluid, retrieved from the hole, and then replaced once again with the drilling tool and
borehole advancement continues to the next sampling depth. For uncased holes, the drilling fluid is
relied upon to stabilize the borehole and prevent it from caving or heaving. In particularly weak or
porous formations where drilling fluids are rapidly lost, cased holes are generally used. In uncased
holes, the drilling fluid is usually recirculated from a mud tank or pit at the ground surface. Borings
that use casing advancers typically use pure water that is not recirculated.

Rotary Drilling Advantages


The advantage of rotary drilling is the relative speed of advancement in deep borings while
maintaining borehole stability that best preserves in-situ soil conditions by counteracting soil and
pore-water pressures in partially or fully saturated conditions. It is of particular advantage in very soft
materials that are very sensitive to disturbance by the drilling equipment. Because of its ability to
maintain natural conditions, rotary drilling is usually the best choice when conducting in-situ analysis
such as vane shear and pressuremeter testing. The trade-offs for rotary drilling is the introduction of
moisture and other minerals that will influence the natural moisture conditions, and the difficulties with
installing groundwater monitoring instruments although this later can in some cases be rectified by
the use of special drilling fluids and by purging the borehole prior to installation. Special care is
needed to contain drilling fluids during exploration, and for ultimate disposal that may involve
transport off-site.

Drill Rods
A variety of drilling rods, casings, and drill bits are available for various tasks. Most drilling tools come
in standard sizes that are generally adaptable to one another. However, complexities arise when
changing from one size to another when various thread sizes and configurations are used. Use the
following information relating to drill rods and casing sizing:

Drill rod and casing sizes are designated from smaller to larger by the letters R, E, A, B, N,
and H. Drill rod outside diameters range from 13/32 inches (27.8mm) for R-sized rods to 3.5
inches (88.9mm) for H-sized rods.

Drill casing outside diameter sizes range from 17/16 inches (36.5mm) for R-sized casing to 4.5
inches (114.3mm) for H-sized casing. Additional letters such as HW or NWJ designate
different thread or coupling configurations.
Complete tables of drilling tool types, sizes, weights, and volumes are available from the
drilling suppliers and manufacturers.

Volume 1
Manual

ODOT Geotechnical Design


36

April 2010

The important aspects of tool size is that the larger diameter, heavier drill sizes generally
provide a more stable hole and allow a greater variety of testing and sampling tools to be
used. These larger sizes also help control the eccentric movement of longer drill strings,
reduce vibration at the drill bit, and help the driller maintain a straight and plumb boring.

The Diamond Core Drill Manufacturers Association (DCDMA) has standardized the drill rod and
casing sizes although any number of other sizes and types remain on the market or are frequently
introduced.

Drill Bits
The choice of drill bit greatly influences the test boring quality and speed of completion. Rotary drill
bits come in a variety of different types, each suited to a particular soil and/or rock composition.
Driller preference is usually what determines what type of bit is used. Experienced drillers can and
should normally be relied upon to select the appropriated bit. Certain drill bits are intended for
specific geologic materials, but many drillers, through their experience and specific equipment, are
able to achieve superb results with bits that are not usually used for that type of material. Follow
these guidelines when using drill bits:

Soft or loose soils: Soft or loose soils are usually drilled with drag bits. These bits have two
or more wings of either tempered steel or carbide inserts that act as cutting teeth.
Hard soils and rock: Roller bits are used to penetrate hard soils and rock. Roller bits may
consist of hardened steel teeth or carbide buttons. Typically, steel teeth are sufficient for
hard soil drilling while carbide button bits are used for bedrock drilling or for drilling in
formations with numerous boulders and potential obstructions.

Rotary Drilling Fluids


Various admixtures are available for mixing with the drilling fluids in different applications. Usually,
the drilling fluid or mud is a mineral solution (usually bentonite and water, thus, a colloidal fluid) with
a viscosity and specific gravity that is greater than water. These properties allow the fluid to better
stabilize the borehole, cool and lubricate the bit, lift the cuttings out of the hole, and can also increase
sample recovery. Various chemical and mineral additives may also be added to the mud mixture for
the site-specific conditions. Certain chemical additives, such as pH stabilizers and flocculants, are
introduced for common groundwater or mineral conditions that are the source of particular drilling
difficulties. Mineral additives, such as barite, may be used to further increase the specific gravity of
the mud for unstable boreholes and zones of high artesian pressures. Other additives inhibit
corrosion of tools; seal off highly fractured or porous formations to prevent fluid loss, increase the
suspension and entrainment of sediments to flush the borehole, and numerous other applications.
Fluids or mud mixtures can greatly enhance rotary drilling, and in some very difficult drilling
situations, is the only way to complete borings. Mud mixing should be treated with care as improper
materials and quantities can actually be detrimental. Volumes and weights should be carefully
measured and fluid density and viscosity should be monitored during borehole advancement as
these properties will be affected by the formation materials. Several batches may be needed for
individual borings depending on the depth of the borehole and other conditions.
The U.S. Bureau of Reclamation and the U.S. Natural Resources Conservation Service have
established general guidelines for drilling mud mixtures including amounts of dry materials, volume of
water, and fluid densities. ASTM D 4380-84 describes the procedures for determining the density of
bentonitic slurries that can be used in rotary drilling.

Volume 1
Manual

ODOT Geotechnical Design


37

April 2010

3.6.2.4 Rock Coring


Rock core sampling is used to obtain a continuous, relatively undisturbed sample of the intact rock
mass for evaluation of its geologic and engineering characteristics. When performed appropriately,
core drilling produces invaluable subsurface information. Rock coring procedures have generally
remained the same since the advent of the technology: a steel tube with a diamond bit rotated into
the rock. Advancements in the bits, core barrels for retrieving the samples, and improvements to
mechanized equipment overall have greatly enhanced this method.
Note:
Rock core drilling procedures and equipment has largely been standardized by
ASTM D2113-99. The Diamond Core Drill Manufacturers Association (DCDMA) has also
standardized bit, core barrel, reaming shell, and casing sizes similar to drill rods.
Rock coring almost exclusively involves the use of diamond bits, thus the terms rock coring and
diamond drilling are used interchangeably. Selecting the proper drill bit for the rock coring
conditions is essential. Sample recovery and drill production is dependent upon it. The ultimate
responsibility for bit selection is the drillers, however, it is important to be familiar with bit types to help
determine recovery problems in the field since they may actually be unrelated to the drilling method.
The actual configuration of the drill bit is selected based on the actual site conditions. The crosssectional configuration, kerf, crown, and number of water ports are all determined by the anticipated
conditions and characteristics of the rock mass. Consider the following:

Incorrect bit selection can be extremely detrimental to core recovery, production, and project
budget.

Typically, a surface-set bit consisting of industrial diamonds set in a hardened matrix is used
for massive rock bodies.

Larger and fewer diamonds in the set are used for soft rocks while smaller and more
numerous diamonds are used in hard rock. Hard rock bits commonly have a rounded or
steeply-angled crown.

Flat-headed bits are usually for very soft rock. Impregnated bits consist of very fine diamonds
in the matrix and are generally used for soft, severely weathered and highly fractured
formations. Some carbide blade and button bits are used for soft, sedimentary rocks. These
are ideally suited for soft rocks with voluminous cuttings that require a considerable amount
bit flushing and cutting extraction.

Core Barrel
The core barrel is the section of the drill string that retains the core specimens and allows them to be
retrieved as a whole section. Core barrels may be of different types and sizes, and may consist of
numerous components that may be changed depending on the rock mass condition. Core barrels
have evolved greatly over time. Single-tube barrels were originally used and required the entire drill
string to be retracted to withdraw the sample. These have evolved through double-tube systems of
either rigid-types where the inner tube rotates with the outer barrel, or swivel-types where the inner
tube remains stationary. Most core barrels used today are triple-tube systems that employ another
non-rotating liner to a swivel-mounted double core barrel. This split metal liner retains the sample
during extraction that allows minimal sample handling and disturbance prior to measurement and
observation. Where desired, a solid, clear plastic tube can be used in place of the split metal tube.
Single and even double-tube coring system often require a considerable amount of effort to extract
the cores from the barrel that can result in detrimental sample disturbance.
Volume 1
Manual

ODOT Geotechnical Design


38

April 2010

Consider the following:

Available triple-tube coring systems usually provide specimens that range in diameter from
15/16 inches (33.5mm) for B-sized core to 39/32 inches (83mm) for P-sized core.

Larger core sizes are also available from rather specialized systems.

A substantial penalty on the quality of rock structural information results from smaller
diameter cores. Most rock core taken is H-sized (213/32 inches, 61.1mm) in diameter.

The use of smaller N-sized cores may be necessary in difficult access, or very deep drilling
applications.

The difference in RQD measurements between single, double, and triple tube systems are
substantial.

Specialized Methods
These specialized methods are also used:

Oriented core barrels: Orienting core barrels can be used to determine the true attitudes of
discontinuities in the rock mass. These specialized core barrels usually scribe a reference
mark on the core as it is drilled. Recording devices within the core barrel relate the known
azimuth to the reference mark so that the exact orientation of the discontinuities can be
determined after the sample has been retrieved.

Borehole camera surveys: Borehole camera surveys are used to determine discontinuity
orientations. Several methods for both oriented coring and down-hole surveying have
evolved, and highly trained personnel are typically needed to operate them successfully. The
1988 AASHTO Manual is a good source of information on the older core orientation systems
while vendors such as the Baker-Hughes Corporation have technical information on the
newer magnetic/electronic core alignment systems.

3 . 6 . 2 . 5 Vib r a t o r y o r S o n i c D r i l l i n g
Sonic drilling may be called vibratory or rotosonic drilling. This type of drilling is used for continuous
sampling in unconsolidated sediments and soft, weathered bedrock. It is best suited for use in
oversized unconsolidated deposits enriched with cobbles and boulders such as talus slopes,
colluvium, and debris flows or any other formation containing large clasts.

Benefits

The primary benefit of this method is recovery of oversized materials in a continuous sample,
rapid drilling rate, reduced volume of cuttings, and fast monitoring well installation.

This drilling technique is 8 to 10 times faster than hollow stem augering and produces about
10% of the volume of cuttings.

Drawbacks

The drawbacks to this method are that it is typically more expensive, and cannot penetrate
very far into bedrock.

The vibration of the drill stem during borehole advancement may disturb the subsurface
materials for an unknown distance ahead of the bit, and soft, loose materials can be liquefied
during sampling.

Volume 1
Manual

ODOT Geotechnical Design


39

April 2010

The sample size and speed of extraction will require additional personnel to process, log, and
classify in the field.

Sonic drill rigs use hydraulic motors that drive eccentric weights to oscillate the drill head. The
oscillation generates a standing sinusoidal wave in the drill stem with a frequency that can be varied
depending on the materials encountered. The drill head also rotates the drill stem. An inner and
outer casing is advanced so that the hole can be cased at the same time that samples are collected.
During drill advancement, the sample is forced into the inner casing from which it is retrieved on a set
interval. SPTs and Shelby tube samples can be taken between runs of rotosonic coring.

3.6.2.6 Becker Hammer Drilling


Becker hammer drills are specifically for use in sand, gravel, and boulders. Some Becker hammer
drill operators may also have a coring system that can also be run for limited applications. Becker
hammer drills use a small diesel-powered pile hammer to drive a special double-walled casing. The
casing can be fitted with an array of toothed bits depending on the application. An air compressor
forces air through the annulus between the casings to the bottom of the hole where it extracts the
materials up through the center of the innermost casing, through a cyclone, and into the sampling
bucket. The materials can be extracted on a set interval as the driller engages the air compressor.
The Becker drill casings range in size from 5.5-inch (14cm) to 9 inches (23cm) for the outer casing,
and 3.3-inch (8.4cm) to 6 inches (15.2cm) respectively for the inner casing. This size of casing
allows retrieval of relatively large, unbroken clasts. As the drill is advanced, blow counts are taken
along with measurements of the hammers bounce chamber pressure. Becker hammer drill data can
be correlated to the soil density and strength in coarse-grained soils similarly to the SPT test. In
addition, SPTs can be taken through the inner casing of the Becker hammer string.

3 . 6 . 2 . 7 S u p p l e m e n t a l D r i l l i n g / E x p l o r a t i o n Ap p l i c a t i o n s
A wide assortment of exploration techniques are available to supplement the subsurface information
gathered from test borings at a project site. Typically, any method that can be employed to properly
evaluate the subsurface conditions in a supplementary capacity is acceptable on an ODOT project if
not constrained by environmental considerations. These methods are usually the most simple and
economic to quickly gather subsurface information with minimal cost. In some cases, more extensive
and costly methods are required to obtain critical design information. Generally, supplemental
investigations consist of simple hand auger borings or backhoe test pits to gather more detailed
information and collect additional samples in near-surface or overburden materials.

Hand Tools
Hand augers are available in many forms that allow rapid penetration of near-surface soils and
collection of representative samples. Various bits can be used that are suited to general soil
conditions that help penetrate and retain samples from certain materials. Extra sections of rods can
be added to extend the depth range of these tools. Small engine-powered augers can also be used
to increase the depth of penetration and to reduce the physical workload. Most hand augers are of
sufficient diameter to permit undisturbed Shelby-tube sampling in the boring where soft soils are
encountered. Additional tools such as jacks, cribbing, and extra weights may be needed to retract
the tube after sampling. Most field vehicles are equipped with shovels that geotechnical designers
can apply to subsurface investigations. Hand-excavated pits can provide essential, detailed
information on the near-surface environment.

Volume 1
Manual

ODOT Geotechnical Design


40

April 2010

Various hand probes and penetrometers can be used to make soundings of soft material depths and
delineate underground facilities in soft ground conditions. Hand auger borings and hand-excavated
test pits are often required for collection of bulk samples.

Cone Penetrometers
Cone penetrometers can be operated from most drill rigs, or they may come as a separate vehicle
specially rigged for cone penetration testing. The cone penetration test (CPT) is conducted by
pushing an instrumented cylindrical steel probe at a constant rate into the subsurface with some type
of hydraulic ram. The cone penetration test is very advantageous in certain (usually soft) soil
conditions as it provides a continuous log of stress, pressures, and other measurements without
actually drilling a hole. CPTs can be conducted with a transducer to measure penetration pore
pressure. Additional instrumentation can be used to measure the propagation of shear waves
generated at the surface. Standard cone penetration test procedures are described in ASTM D
3441-98. Electronic CPT testing must be done in accordance with ASTM D 5778.

Percussion or Direct push (i.e. GeoProbe) Borings


Direct push drills are hydraulically-powered, percussion/probing machines originally intended for use
in environmental investigations. The direct push method uses the weight of the vehicle combined
with percussion to advance the drill string. Drive tools are used to obtain continuous, small-diameter
soil cores or discrete samples from specific locations. Direct push drills can obtain continuous
samples through the soil column and are capable of penetrating most soils up to about 100 feet
(30m). Small-diameter piezometers can also be installed through the direct push tools. Direct push
rigs are quick and economical to mobilize and sample the soil column very quickly. Their small
diameter and method of penetration produce few if any cuttings that must be disposed of. The
percussion advance of the direct push method produces a considerable amount of sample
disturbance.
Note:
Direct push advancement rates may provide a relative determination of soil density with respect to
material encountered by that particular machine but it is not correlative to SPT data. Direct push rigs
are lighter and less powerful than most conventional drill rigs. Thus, they do not have the ability to
penetrate certain formations, and because of the effort in doing so, may give a false, overestimation
of the formation density.

Test Pits
Backhoe-excavated test pits or trenches are commonly used to provide detailed examination of near
surface geologic conditions and to collect bulk samples. Test pits allow examination of larger-scale
features that would not be visible in standard borehole samples. Features such as faulting, seepage
zones, material contact geometry and others are readily measured in test pit walls. In addition,
Torvane and pocket penetrometer tests can be performed in the walls and floor of the test pit. Inplace percolation testing can also be carried out in test pits. Test pits have the advantage of the
shear bulk of materials that can be observed. In this regard, the overall composition of the materials
in a unit are better assessed by the many cubic feet of material excavated and observed opposed to
the relatively minute amount of material contained in a split spoon sampler.

Volume 1
Manual

ODOT Geotechnical Design


41

April 2010

Warning:
Under no circumstances will personnel enter a test pit deeper than 4 feet (1.2m) below the ground
surface unless the appropriate shoring and bracing is used. If any evidence of instability or seepage
is evident in the test pit walls, no entry will be permitted until shoring is complete. Test pits must be
filled in as soon as they are completed to prevent passersby from entering or falling in. When a test
pit is used for percolation tests or for assessment of trench stability, appropriate barricades and signs
must be placed around the site to prevent accidental entry.

ODEX or Air-Track Drilling


Percussive air drilling is typically used in a similar manner to other probing systems with the
exception that air-drill holes are used to probe harder materials. A relative rate of advancement
coupled with the cuttings retrieved in certain intervals allows basic interpretation of subsurface
conditions. ODEX systems using an outer casing allow installation of instruments below the water
table that would otherwise be impossible to install with other air-driven equipment. The advantage of
this method is the speed of installation and borehole advancement. As previously described, air
drilling system are not suited for standard testing methods due to the unknown amount of down-hole
disturbance.

3.6.3

Alternative Exploration Methods and Geophysical


Surveys

Alternatives to drilling and test pit excavations characteristically involve the use of geophysical
methods. For ODOT projects, geophysical survey results are always supplemental to direct
observation of subsurface conditions by borings and test pits and should never be considered as a
replacement.
Geophysical surveys play an important role in engineering geology and geotechnical engineering
however they do not provide all of the information needed for the development of geotechnical
design parameters.
Note:
From a liability and construction claims standpoint, direct observation, sampling, and testing are
critical. Direct observation and measurement will assure that subsurface conditions not measured by
geophysical survey methods are revealed and further support or refute the results of geophysical
surveys.
Most of the data obtained from a geophysical survey require an experienced and highly-trained
geophysicist to interpret and process before it is of any use to an engineering geologist or
geotechnical engineer. Geophysicists can base their interpretation on direct calculations, tabulations,
or regression analyses, or they may base it wholly upon their own experience. Any geophysical
method used has its own aspects that can result in serious misinterpretation or inappropriate use of
the results. Prior knowledge of the actual site conditions and the possible errors of the survey
technique are needed to calibrate, or fit the data to the known baseline data.
Geophysical survey results and resolution of the data is dependent upon the density of measurement
points, and frequency of measurements. These variables may be set according to the overall project
needs and level of detail required. Modern geophysical instruments are sensitive enough to produce
measurements at the levels needed for geotechnical investigations. Methods most frequently used
are:

Volume 1
Manual

ODOT Geotechnical Design


42

April 2010

Seismic methods are the most commonly conducted techniques for engineering geologic
investigations.

Seismic refraction provides the most basic geologic data by using the simplest procedures,
and commonly available equipment. The data provided is the most readily interpreted and
correlated to other known material properties.

3.7.1

Geotechnical Instrumentation
General Instrumentation and Monitoring

Of equal importance to site characterization and exploration as sampling and testing data is the
information provided by geotechnical instrumentation and monitoring. Sampling and testing of
materials provides needed design information concerning the existing site conditions at the time of
investigation. Information regarding certain site conditions as they change through time due to the
effects of natural variations in the earths surface and atmosphere or the effects of human activities,
such as construction, can be provided by the appropriate selection, installation, and monitoring of
geotechnical instruments. Most geotechnical instruments are used to monitor the performance of
structures and earthworks during construction and operation of the facility. Some instrumentation
programs are planned to provide actual design criteria such as landslide depths of movement and
piezometric surfaces. Other programs are intended to verify design assumptions. In any case,
considerable design and planning efforts are needed to derive the needed results. Geotechnical
instrumentation has become much more user-friendly as technologies have developed, but an allinclusive process beginning with a determination of the instrumentation project objectives that are
carried through to completion and use of the data.

3.7.2

Purposes of Geotechnical Instrumentation

A rule of thumb for geotechnical instrumentation programs is: every instrument installed should be
selected and placed to assist in answering a specific question. The point of this rule is to start a
geotechnical instrumentation program on the correct course of study to acquire the necessary results
with the greatest efficiency. Instruments can have an initially high installation cost, but the time and
effort for reading them and making sense of the results is where the most costly inefficiencies occur.
Any instrument installed will provide some information; whether or not it is relevant to the immediate
project requirements is the issue. Therefore, efforts must be concentrated on the primary questions
to gather the most important data from the instrumentation program without time lost to the analysis
of extraneous data.

3 . 7 . 2 . 1 S i t e I n ve s t i g a t i o n a n d E x p l o r a t i o n
Instruments are regularly used to characterize the initial site conditions during the design phase of a
project. Landslide remediation projects rely on instruments to determine depths and rates of
movement as well as pore water pressures to provide basic information for stability analysis and
mitigation design.
Most project sites require some information concerning the actual depth and seasonal fluctuation of
groundwater that not only affects the project design, but also its constructability.

Volume 1
Manual

ODOT Geotechnical Design


43

April 2010

3 . 7 . 2 . 2 D e s i g n Ver i f i c at i o n
Instruments are frequently used to verify design assumptions and to check that facility performance is
as expected. Instrument data gathered early in a project can be used to modify the design in later
phases. Geotechnical instruments are also an inherent part of proof testing to verify design
adequacy.

3.7.2.3 Construction and Quality Control


Geotechnical instruments are commonly used to monitor the effects of construction. Construction
procedures and schedules can be modified based on actual behavior of the project features for
ensuring safety as well as gaining efficiency in the actual construction as determinations can be
made regarding how fast construction can proceed without the risk of failure or unacceptable
deflections. Instruments can be used to monitor contractor performance to assure that contract
requirements and specifications are being met.

3.7.2.4 Safety and Legal Protection


Instruments can be used to provide early warning of impending failures allowing time to isolate the
problems and begin implementation of remedial actions. Instrument data provides crucial evidence
for legal defense of the agency should owners of adjacent properties claim that construction or
operations have caused damage.

3.7.2.5 Performance
Instruments are used for the short and long-term service performance of various facilities.
Deformation, slope movement, and piezometric surface measurements in landslides can be used to
evaluate the performance of drainage systems installed to stabilize the landslide. Loads on rock
bolts and tiebacks may be monitored to assess their long-term performance or evaluate the need for
additional supports.

3.7.3

Criteria for Selecting Instruments

For each project, the critical parameters must be identified by the designer that will require
instrumentation to determine. The appropriate instruments should then be selected to measure them
based on the required range, resolution, and precision of measurements. The ground conditions are
another consideration in the choice of instruments. Use the following to help select instruments:

Landslides: Relatively fast-moving landslides may require a larger-diameter inclinometer


pipe or TDR cable to determine the zone of slide movement, or Vibrating Wire piezometers
may be selected to measure groundwater in low permeability soils where a standpipe would
require a large volume of water to flow into it before even small changes in pore-water
pressure can be detected.

Temperature and humidity: Temperature and humidity also affect the choice of instruments.
Certain instruments may be difficult to use in freezing conditions while warm and humid
environments may affect the reliability of electronic instruments unless particular care is taken
to isolate their environment.

Number of parameters: The number of parameters to measure is also important for


instrument selection since soil and rock masses typically have more than one property that
dictates their behavior. Some parameters correlate with one another, and instruments that

Volume 1
Manual

ODOT Geotechnical Design


44

April 2010

obtain complementary measurements provide an efficiency gain. In areas with complex


problems, several parameters can be measured, and a number of correlations can be found
from instrumentation data leading to a better understanding of the site conditions. Strain
gages and load cells on a retaining wall and inclinometers behind it are examples where
complementary data can be obtained. When relationships can be developed with the data,
further data can be obtained even when one set of instruments fail.

Instrument performance and reliability: Instrument performance and reliability are also
important considerations. The cost of an instrument generally increases with higher
resolution, accuracy, and precision in the instrument. Also, the range of measurements
obtained can be reduced by higher-functioning instruments, so the geotechnical designer
should have a clear understanding of the scale and level of measurements to be taken.
Example: An example is the placement of a vibrating wire transducer in a borehole to
measure an unknown piezometric surface. The instrument selected would have a wide
range of testing, but a lower resolution of values that could be read. Where the piezometric
surface is known within a narrower range and small changes are of significance to the
design, an instrument capable of reading a smaller range of values but at a higher resolution
within the known range.

Quality of the instrument: There are some instances where the use of lower-quality
instruments is warranted, but in general, choosing a lower-quality instrument to save on initial
costs is a false economy. The difference in cost between a high-quality instrument and a
lower-quality instrument is low with respect to the overall cost of installing and monitoring an
instrument.

Cost: The cost of drilling a hole and the labor of installing the instrument is usually an order of
magnitude higher than the cost of the instrument. The less easily quantifiable loss of data
from a failed instrument in terms of monetary cost should also be considered. It is expensive
and often impossible to replace failed instruments. Furthermore, essential baseline data is
also lost that cannot be replaced.

3 . 7 . 3 . 1 Au t o m a t i c D a t a Ac q u i s i t i o n S ys t e m s ( AD AS )
Automatic Data Acquisition Systems (ADAS) can provide significant advantages to a geotechnical
instrumentation program. They can provide numerous readings at set and reliable intervals, and they
can store and transmit data from remote or difficult access locations. ADAS are necessary for realtime instrument monitoring and relay. They are beneficial at sites where many sensors are present
that would require copious staff time to read manually or for large-scale proof tests with many
concurrently-read instruments to be monitored throughout the test.
Automatic Data Acquisition Systems come in many forms ranging from the very simple, user-friendly
devices to systems requiring significant programming and electronics to install and run. Project
requirements usually dictate what system is selected, but the simplest, most inexpensive, and easiest
to connect to the chosen instruments are best. Follow these guidelines:

Simple dataloggers connected to individual instruments that are retrieved and downloaded
periodically are sufficient for most projects.

Large, complex problems may require a more intelligent system that can be programmed to
change monitoring routines in response to site or environmental changes.

Volume 1
Manual

ODOT Geotechnical Design


45

April 2010

Most instrumentation companies also have companion dataloggers to go with their products
while several independent companies also manufacture easy-to-use dataloggers. Other
companies, such as Campbell Scientific Incorporated, produce more complex systems that
can read multiple installations of different types of instruments as well as store and transmit
data.

In addition to the data collection devices, these firms also produce software for processing
and displaying the data. The software is another consideration if export to other systems is
desired. Compatibility between programs can create problems and errors in the end product
of an instrumentation project.

3 . 7 . 3 . 2 I n s t r u m e n t U s e a n d I n s t al l a t i o n
Instruments have been developed to monitor many specific geologic conditions and engineering
parameters. In many cases, a single instrument can be used or adapted for use on other
applications. For this, the manufacturer and other professionals should be consulted to assure that
the results obtained are valid, or, they may have insights and case histories that are of use for the
situation. The manufacturers literature, installation procedures, and other guidance documents
should be followed for proper installation of their products as procedures can vary for different
manufacturers same instrument products. Detailed discussions of instrument installation and
initialization procedures, function, and operation can be found in manufacturers documents such as
Slope Indicator Company (SINCO) Applications Guide or in published literature such as Dunnicliff
(1988).

3.7.3.3 Inclinometers
Inclinometers are used on transportation projects mainly to detect and monitor lateral earth
movements in landslides and embankments. They are also used to monitor deflections in laterally
loaded piles and retaining walls. Horizontally installed inclinometers can also be used to monitor
settlement. Inclinometer systems are composed of:

grooved casing installed in a borehole, embedded in a fill or concrete, or attached to


structures,

probe and cable for taking measurements at set intervals in the casing, and

a digital readout unit and/or data storage device.

The installed casing is for single installation use, and the probe, cable and data storage unit are used
for almost all installations.
Note:
It is important to use the same probe for each reading in any particular installation since each probe
must be independently calibrated.
Inclinometers are manually read by a trained technician on a set schedule or in response to
environmental changes such as increased rainfall in the area or observation of surficial signs of slope
movement. In-place inclinometers spanning known or highly suspected zones of movement can be
installed for continuous, automatic monitoring. These usually remain in the hole permanently if
significant slope movement occurs.

Volume 1
Manual

ODOT Geotechnical Design


46

April 2010

Inclinometer casing installation is essential to successful performance of the instrument.


Shortcuts taken during installation will frequently result in poor performance of the instrument
or render it completely useless.

Inclinometers should be installed according to the procedures described in the SINCO


Applications guide with the exception of the grout valve.

Borings should be initially drilled or later reamed to a sufficient diameter that will
accommodate the inclinometer casing and an attached tremie tube.

The tremie tube should be attached to the inclinometer casing approximately 6 inches above
the bottom and along the casing at a close enough interval to prevent it from getting tangled
or constricted in the borehole.

One of the four grooves in the inclinometer casing should be aligned to the direction of slide
movement as the casing is assembled and lowered into the hole to prevent spiraling.

If the borehole walls are unstable, the drill casing may need to remain in the borehole, and
withdrawn as the grout level rises. Generally, the grout should be maintained at a visible level
in the casing as the drill string is withdrawn.

Initial readings should be taken as soon as the grout has sufficiently set up. This is usually 3 to 5
days after grouting. During installation, some grout is naturally lost to fractures and voids in the
formation. This may occur to the extent that additional grouting is required. Usually, this only entails
topping off the hole with a small batch of grout to stabilize the uppermost portion of the casing. In
more severe cases, the grout pump may be reconnected to the tremie tube to re-grout the remaining
voids.

3.7.3.4 Piezometers
Piezometers used to measure pore-water pressure and groundwater levels can range from simple
standpipes to complex electronic devices or pneumatic systems. Piezometers are typically installed
in selected layers to measure the piezometric pressures in that layer. The layout and target depths of
piezometer installation are determined by actual site conditions and project requirements.
Note:
All piezometers must be installed according to Oregon Water Resources Department regulations
defined by ORS 690.240 and ORS 537.747 through ORS 737.799 (appropriation of water
generally). Specifications for a properly operating instrument are usually more stringent than these
rules apart from the requirements for abandonment.
The various types of piezometers are generally used for different applications as described below.

Standpipe piezometers are general-purpose instrument for monitoring piezometric water


levels and are best-suited for granular materials. Standpipe piezometers require a water
level indicator to obtain readings.

Vibrating Wire piezometers utilize a pressure transducer to convert water pressure to a


frequency signal that is read by an electronic device. Vibrating Wire piezometers can be
automated by electronic systems.

Pneumatic piezometers are typically used to measure pore water pressure in saturated
conditions. Both Pneumatic and vibrating wire piezometers are used for all soil types and are

Volume 1
Manual

ODOT Geotechnical Design


47

April 2010

better suited to fine-grained soils than the standpipe variety due to the response time and
volume of water needed to record changes in water level in that type.
Piezometers should be placed at the desired sensing zone in a porous medium and sealed with the
appropriate materials above and below this zone to assure measurement of the piezometric pressure
in the desired location. Porous mediums or filter packs should be composed of pre-screened
commercial-grade silica sand. All piezometers should be installed and initialized according to their
manufacturers specifications.

3.7.3.5 Other Instruments


A vast array of geotechnical instruments is available for most applications. Strain gauges,
extensometers and load cells of all types and configurations for structural as well as geotechnical
applications are obtainable from numerous vendors. Most vendors have prescribed applications as
well as installation and monitoring procedures that should be followed when using their products on
transportation projects. Professional knowledge, experience, and judgment must be applied to the
use of all instruments to assure appropriate use of these instruments and the adequacy of data
obtained.

Environmental
Exploration

Protection

during

Compliance with all State, Federal, and Local ordinances, laws and regulations concerning
environmental protection at all work locations is mandatory for any activity that may disturb the
ground surface or vegetation. All environmental permits, clearances, or any other documentation
needed for compliance with the pertinent environmental regulations must be ready prior to
mobilization of exploration equipment.
The ODOT Programmatic Biological Opinion for Drilling, Surveying, and Hydraulic Engineering
Activities may be applicable for some sites. This document can be referenced on the ODOT GeoEnvironmental web page.
Note:
Every precaution necessary to minimize environmental impacts during site investigation must be
taken, and every effort made to restore the site to its original condition. All drilling fluids and cuttings
must be disposed of safely and legally. In no circumstance should sediment-laden water or other
pollutants be allowed to enter streams or other bodies of water. In the event where there is a
potential for pollutants to contaminate such, all operations will be suspended until the situation can be
rectified. Violation of Federal, State, and Local environmental protection laws can result in personal
penalties, including arrest and incarceration.

3.8.1

Protection of Fish, Wildlife, and Vegetation

Compliance with the Laws of the Oregon Department of Fish and Wildlife, National Marine
Fisheries Service, United States Fish and Wildlife Service, and the rules and practices
developed through the Oregon Plan for Salmon and Watersheds is also mandatory. All
subsurface investigation activities shall be conducted to avoid any hazard to the safety and
propagation of fish and shellfish in the waters of the State.
Unless specifically authorized by the State and by permit, the Contractor shall not:

Volume 1
Manual

ODOT Geotechnical Design


48

April 2010

Use water jetting

Release petroleum or other chemicals into the water, or where they may eventually enter the
water

Disturb spawning beds or other wildlife habitat

Obstruct streams

Cause silting or sedimentation of water

Use chemically treated timbers or platforms

Impede fish passage


The permitted work area boundaries will be defined by the permit for the project from the regulatory
agencies.

3.8.2

Forestry Protection

All necessary permits must be obtained prior to exploration in accordance with ORS 477.625 and
ORD 527.670, and comply with the laws of any authority having jurisdiction for protection of forests.
At certain times of the year, the exploration activities will be subject to IFPL constraints, and
operational schedules must be adjusted accordingly. Fire-suppression equipment may be required
on site as well as a designated fire watch.

3.8.3

Wetland Protection

All operations shall comply with the Clean Water Act Section 404 (33 U.S.C. 1344); Federal Rivers
and Harbors Act of 1899, Section 10 (33 U.S.C. 403 et seq.); Oregon Removal-Fill law (ORS
196.800 - 196.990); Oregon Removal and Filling in Scenic Waterways law (ORS 390.805 390.925), and other applicable Laws governing preservation of wetland resources.
Note:
The terms wetland, or wetlands are defined as Areas that are inundated or saturated by surface
or groundwater at a frequency and duration sufficient to support, and that under normal circumstance
do support, vegetation typically adapted for life in saturated Soil conditions. Wetlands generally
include swamps, marshes, bogs, and similar areas. Wetlands also include all other jurisdictional
waters of the U.S. and/or the State.
If wetlands are known to be on the project site, they should be delineated by the regions wetland
specialist or their contractor to prevent accidental entry by the exploration operation. Wetlands to be
temporarily impacted should also be identified at this time. Wetlands to be protected will be
considered as no work zones.
Subsurface exploration operations must also comply with Clean Water Act Section 404 permits
issued by the U.S. Army Corps of Engineers, and Fill/Removal permits issued by DSL. These permits
allow specified quantities of fill and excavation, including soil and rock samples within specifically
identified areas of wetlands.

3.8.4

Cultural Resources Protection

The exploration crew is also required to comply with all Laws governing preservation of cultural
resources. Cultural resources may include, but are not limited to, dwellings, bridges, trails, fossils,
and artifacts. Known locations of cultural resources will be considered as no work zones.
Volume 1
Manual

ODOT Geotechnical Design


49

April 2010

If cultural resources are encountered in the project area, and their disposition is not addressed in the
contract, the exploration crew shall:

Immediately cease operations or move to another area of the project site

Protect the cultural resource from disturbance or damage

Notify the regions cultural resource specialist

The regions cultural resource specialist will:

Arrange for immediate investigation

Arrange for disposition of the cultural resources

Notify the exploration crew when to begin or resume operations in the affected area

Volume 1
Manual

ODOT Geotechnical Design


50

April 2010

REFERENCES

AASHTO, 2007, LRFD Bridge Design Specifications, American Association of State Transportation
and Highway Officials, 17th Edition (with current Interims), Washington, D.C., USA.
American Association of State Highway and Transportation Officials, Inc., 1988, Manual on
Subsurface Investigations.
C.H. Dowding, Ed., Site Characterization & Exploration, ASCE Specialty Workshop Proceedings,
Northwestern University, 1978.
Dunnicliff, John 1988. Geotechnical Instrumentation For Monitoring Field Performance, John Wiley &
Sons, New York.
U.S. Department of Transportation, Federal Highway Administration Evaluation of Soil and Rock
Properties, Geotechnical Engineering Circular No. 5, FHWA-IF-02-034, April, 2002.
U.S. Department of Transportation, Federal Highway Administration Subsurface Investigation
Participant's Manual, Publication No. FHWA HI-97-201, November, 1997.
U.S. Department of Transportation, Federal Highway Administration Subsurface Investigations Geotechnical Site Characterization Reference Manual, Publication No. FHWA NHI-01-031, May,
2002.
Turner, Keith A., and Schuster, Robert L., Eds., LANDSLIDES Investigation and Mitigation,
Transportation Research Board Special Report 247, 1996, Pages 140-163.
Geophysical Exploration for Engineering and Environmental Investigations, U.S. Army Corps of
Engineers Engineering and Design Manual, EM 1110-1-1802, August 1995.
Geotechnical Investigations, U.S. Army Corps of Engineers Engineering and Design Manual, EM
1110-1-1804, January 2001.
U.S. Department of the Interior, Bureau of Reclamation, 1994, Engineering Geology Field Manual.
References are made to various ASTM standards. The ASTM International standards located at
www.astm.org (the ASTM Site).

Volume 1
Manual

ODOT Geotechnical Design


51

April 2010

Appendix 3-A Permit of Entry Form


Oregon Department of Transportation

RIGHT-OF-ENTRY for EXPLORATION


REGION 3 GEOLOGY
Phone: (541) 957-3602 FAX: (541) 957-3604
3500 NW Stewart Parkway
Roseburg, OR 97470
(1) (We) ______________ and __________________________ hereinafter referred to as
grantor, do hereby grant to the STATE OF OREGON, by and through the Oregon
Department of Transportation, and its officers, agents, and employees, the right and
license to go upon the following described real property to drill or to gain access to
highway Right-of -Way for exploration core drilling at:
Township 37 South, Range 2 West, Section 28
77 Hanley Road
Central Point, Oregon 97502
Property Description:
D-89-16328
37-2W-28 TL 800
IT IS UNDERSTOOD AND AGREED: That this right and license shall be valid until all
exploration is completed unless revoked by grantor before completion. It is further
understood that the Oregon Department of Transportation shall, to the extent permitted by
Oregon law, be responsible for any unnecessary damage done, in connection with said
exploration, this will include any crops or other improvements on said property.
Grantor hereby represents and warrants that He/She is the owner of said property or
otherwise has the right to grant this permit of entry.
Date____________ Day________, 2003
Permission Acquired by: _____________________
Signature:__________________________________

Volume 1
Manual

ODOT Geotechnical Design


52

April 2010

Title: Project Geologist


Owner(s)
Signature(s):___________________________________________________________

Volume 1
Manual

ODOT Geotechnical Design


53

April 2010

Appendix 3-B Utility Notification Worksheet


Memo to File
UTILITY LOCATE DATA SHEET
Region Geology Unit
Oregon Department of Transportation
Project Name:
Highway and Mile Point:
Utility Locate Called By:
Locators Called (When):

Required Information
Caller ID #:
Type of Work:
County/City
Highway:
Mile Point:
Township/Range/
Quarter Section:
Distance from
Nearest Cross
Street:
Overhead Lines:
Special Markings:
Date to Be Located:
Ticket#:
Name of Person Called:
Utilities Notified:
Utilities Field Marked:
Gas
Electric
Sewer
Water
Telephone
Cable Television
Irrigation
Signals/Illumination

Volume 1
Manual

ODOT Geotechnical Design


54

April 2010

Other

Volume 1
Manual

ODOT Geotechnical Design


55

April 2010

Chapter

Soil and Rock Classification and


Logging

General

The ODOT Soil and Rock Classification Manual (1987) should be used for the description and
classification of all soil and rock materials. This manual is available on the Geo-Environmental web
page at the following address:
Soil_Rock_Classification_Manual.pdf on ftp.odot.state.or.us

Volume 1
Manual

ODOT Geotechnical Design


1

April 2010

Volume 1
Manual

ODOT Geotechnical Design


2

April 2010

Chapter

Engineering Properties of Soil and


Rock

General

The purpose of this chapter is to identify appropriate methods of soil and rock property assessment
and describe how to use soil and rock property data to establish engineering parameters for
geotechnical design. Soil and rock design parameters should be based on the results of a
geotechnical investigation which includes in-situ field testing and a laboratory testing program, used
separately or in combination. The geotechnical designers responsibility is to determine which
parameters are critical to the design of the project and then determine the parameters to an
acceptable level of accuracy. See Chapter 2 and the individual chapters that cover each
geotechnical design element area for further information on how to plan and obtain soil and rock
parameters.
The detailed measurement and interpretation of soil and rock properties should be consistent with the
guidelines provided in Sabatini, et al, April, 2002, U.S. Department of Transportation, Federal
Highway Administration Evaluation of Soil and Rock Properties, Geotechnical Engineering
Circular No. 5, FHWA-IF-02-034.
The focus of geotechnical design property assessment and final selection should be on the individual
geologic strata identified at the project site. A geologic stratum is characterized as having the same
geologic depositional history and stress history, and generally has similarities throughout the stratum
in terms of density, source material, stress history, and hydrogeology. It should be recognized that the
properties of a given geologic stratum at a project site are likely to vary significantly from point to point
within the stratum. In some cases, a measured property value may be closer in magnitude to the
measured property value in an adjacent geologic stratum than to the measured properties at another
point within the same stratum. However, soil and rock properties for design should not be averaged
across multiple strata. It should also be recognized that some properties (e.g., undrained shear
strength in normally consolidated clays) may vary as a predictable function of a stratum dimension
(e.g., depth below the top of the stratum). Where the property within the stratum varies in this
Volume 1
Manual

ODOT Geotechnical Design


1

April 2010

manner, the design parameters should be developed taking this variation into account, which may
result in multiple values of the property within the stratum as a function of a stratum dimension such
as depth.

Volume 1
Manual

ODOT Geotechnical Design


2

April 2010

Influence of Existing and Future


Conditions on Soil and Rock Properties

Many soil properties used for design are not intrinsic to the soil type, but vary depending on
conditions. In-situ stresses, the presence of water, rate and direction of loading can all affect the
behavior of soils. Prior to evaluating the properties of a given soil, it is important to determine the
existing conditions as well as how conditions may change over the life of the project. Future
construction, such as new embankments, may place new surcharge loads on the soil profile or the
groundwater table could be raised or lowered. Often it is necessary to determine how subsurface
conditions or even the materials themselves will change over the design life of the project. Normally,
consolidated clays can gain strength with increases in effective stress and over-consolidated clays
may lose strength with time when exposed in cuts. Some construction materials such as weak rock
may loose strength due to weathering within the design life of the embankment.

Methods of Determining Soil and Rock


Properties

Subsurface soil or rock properties are generally determined using one or more of the following
methods:

in-situ testing during the field exploration program,

laboratory testing, and

back analysis based on site performance data.

The two most common in-situ test methods for use in soil are the Standard Penetration Test, (SPT)
and the Cone Penetrometer Test (CPT). Other in-situ tests, such as pressuremeter and vane shear
are used less frequently, but are important tests in specific instances. In-situ tests for rock are
sometimes performed for the design of major structures but generally are not common for highway
applications.
The laboratory soil and rock testing program generally consists of index tests to obtain general
information or to use with correlations to estimate design properties, and performance tests to directly
measure specific engineering properties. A wide array of index and performance tests for soil, rock
and groundwater measurement are discussed in Sabatini, et al, April, 2002, U.S. Department of
Transportation, Federal Highway Administration Evaluation of Soil and Rock Properties,
Geotechnical Engineering Circular No. 5, FHWA-IF-02-034.
The observational method, or use of back analysis, to determine engineering properties of soil or
rock is often used with slope failures, embankment settlement or excessive settlement of existing
structures.

Landslides or slope failures: With landslides or slope failures, the process generally starts
with determining the geometry of the failure and then determining the soil/rock parameters or
subsurface conditions that cause the safety factor to approach 1.0. Often the determination of
the back-calculated properties is aided by correlations with index tests or experience on other
projects.

Volume 1
Manual

ODOT Geotechnical Design


3

April 2010

Embankment settlement: For embankment settlement, a range of soil properties is


generally determined based on laboratory performance testing on undisturbed samples.
Monitoring of fill settlement and pore pressure in the soil during construction allows the soil
properties and prediction of the rate of future settlement to be refined.

Structure settlement: For structures such as bridges that experience unacceptable


settlement or retaining walls that have excessive deflection, the engineering properties of the
soils can sometimes be determined if the magnitudes of the loads are known. As with slope
stability analysis, the geometry of the subsurface soil must be adequately known, including
the history of the groundwater level at the site.

In-Situ Field Testing

Standards and details regarding field tests such as the Standard Penetration Test (SPT), the Cone
Penetrometer Test (CPT), the vane shear test, and other tests and their applications in geotechnical
design are provided in Sabatini, et al. (2002). Standards for sampling and testing of materials are in
general accordance with ASTM (www.astm.org).
In general, correlations between N-values and soil properties should only be used for cohesionless
soils and sand, in particular. Caution should be used when using N-values obtained in gravelly soil.
Gravel particles can plug the sampler, resulting in higher blow counts and estimates of friction angles
than actually exist. Caution should also be used when using N-values to determine silt or clay
parameters due to the dynamic nature of the test and resulting rapid changes in pore pressures and
disturbance within the deposit. Correlations of N-values with cohesive soil properties should generally
be considered as preliminary. N-values can also be used for liquefaction analysis. See Chapter 6 for
more information regarding the use of N-values for liquefaction analysis.
A discussion of field measurement of permeability is presented in Sabatini, et al. (2002), and ASTM D
4043 presents a guide for the selection of various field methods.
Note:
If in-situ test methods are utilized to determine hydraulic conductivity, one or more of the following
methods should be used:

Well pumping tests

Packer permeability tests

Seepage Tests

Slug tests

Piezocone tests

5.4.1

Correction of Field SPT Values

The N-values obtained are dependent on the equipment used and the skill of the operator, and
should be corrected to standard N60 values (an efficiency of 60 percent is typical for rope and
cathead systems). This correction is necessary because many of the correlations developed to
determine soil properties are based on N60-values. SPT N-values should be corrected for hammer
efficiency in accordance with section 4.4.3 of Sabatini, et al. (2002).

Volume 1
Manual

ODOT Geotechnical Design


4

April 2010

ODOT requires that all hammers have an energy measurement performed at the time of drilling of a
boring or that the hammer efficiency of each hammer be supplied with the boring log. Caution must
be used when noting N-value correlations and the notation N(uncorr) or N(60) must be indicated.
The following values for energy ratios (ER) may be assumed if hammer specific data are not
available:

ER = 60% for conventional drop hammer using rope and cathead

ER = 80% for automatic trip hammer

Hammer efficiency (ER) for specific hammer systems used in local practice may be used in lieu of
the values provided. If used, specific hammer system efficiencies shall be developed in general
accordance with ASTM D-4945 for dynamic analysis of driven piles or another accepted procedure.
Corrections for rod length, hole size, and use of a liner may also be made, if appropriate. In general,
these are only significant in unusual cases or where there is significant variation from standard
procedures. These corrections may be significant for evaluation of liquefaction. Information on these
additional corrections may be found in: Proceedings of the NCEER Workshop on Evaluation of
Liquefaction Resistance of Soils; Publication Number: MCEER-97-0022; T.L. Youd, I.M. Idriss (1997)
and in Cetin, K., Seed, R., et al.
N-values are also affected by overburden pressure, and in general should be corrected for that effect,
if applicable to the design method or correlation being used. Corrections for overburden pressure are
normally only applied to cohesionless soils where the resulting N-values will be used in correlation
with the angle of internal friction or liquefaction analysis or to obtain other design parameters. Nvalues corrected for both overburden and the efficiency of the field procedures used shall be
designated as (N1)60 as stated in Sabatini, et al. (2002).

Laboratory Testing of Soil and Rock

Laboratory testing is a fundamental element of a geotechnical investigation. The ultimate purpose of


laboratory testing is to measure physical soil and rock properties utilizing standard repeatable
procedures. Laboratory test data is also used to refine the visual observations and field testing data
from the subsurface field exploration program, and to determine how the soil or rock will behave
under the proposed loading conditions. The ideal laboratory program will provide sufficient data to
complete an economical design without incurring excessive tests and costs. Depending on the
project issues, testing may range from simple soil classification testing to complex strength and
deformation testing. Details regarding specific types of laboratory tests and their use are provided in
Sabatini, et al. (2002).

5.5.1

Quality Control for Laboratory Testing

Improper storage, transportation and handling of samples can significantly alter the material
properties and result in misleading test results. The requirements provided in FHWA-HI-97021,Subsurface Investigations, NHI course manual #132031, Mayne, et al., (1997) for these issues
and laboratory testing of soils should be followed. Laboratories conducting geotechnical testing shall
be either AASHTO accredited or fulfill the requirements of AASHTO R18 for qualifying testers and
calibrating/verifications of testing equipment for those tests being performed.

Volume 1
Manual

ODOT Geotechnical Design


5

April 2010

5.5.2

Developing the Testing Plan

The amount of laboratory testing required for a project will vary depending on availability of
preexisting data, the character of the soils and the requirements of the project. Laboratory tests
should be selected to provide the desired and necessary data as economically as possible.
Geotechnical information requirements are provided in Sabatini, et al. (2002) that address design of
geotechnical features. Laboratory testing should be performed on both representative and critical test
specimens obtained from geologic layers across the site. Critical areas correspond to locations
where the results of the laboratory tests could result in a significant change in the proposed design. In
general, a few carefully conducted tests on samples selected to cover the range of soil properties
with the results correlated by classification and index tests is the most efficient use of resources. The
following should be considered when developing a testing program:

Project type (bridge, embankment, rehabilitation, buildings, etc.)

Size of the project

Loads to be imposed on the foundation soils

Types of loads (i.e., static, dynamic, etc.)

Whether long-term conditions or short-term conditions are in view

Critical tolerances for the project (e.g., settlement limitations)

Vertical and horizontal variations in the soil profile as determined from boring logs and visual
identification of soil types in the laboratory

Known or suspected peculiarities of soils at the project location (i.e., swelling soils, collapsible
soils, organics, etc.)

Presence of visually observed intrusions, slickensides, fissures, concretions, etc in sample


how will it affect results

Project schedules and budgets

5.6.1

Engineering Properties of Soil


Laboratory Performance Testing

Laboratory performance testing of soil is mainly used to estimate strength, compressibility, and
permeability characteristics. Shear strength may be determined on either undisturbed specimens of
fine- grained soil (undisturbed specimens of granular soils are very difficult, if not impossible, to
obtain), or disturbed or remolded specimens of fine or coarse grained soil. There are a variety of
shear strength tests that can be conducted, and the specific type of test selected depends on the
specific application. See Sabatini, et al. (2002) for specific guidance on the types of shear strength
tests needed for various applications.

5.6.1.1

D i s t u r b e d S h e a r S t r e n g t h Test i n g

Disturbed soil shear strength testing is less commonly performed, and is primarily used as
supplementary information when performing back-analysis of existing slopes, or for fill material and
construction quality assurance when minimum shear strength is required. It is difficult to obtain
Volume 1
Manual

ODOT Geotechnical Design


6

April 2010

accurate shear strength values through shear strength testing of disturbed (remolded) specimens
since the in-situ density and soil structure is quite difficult to accurately recreate, especially
considering the specific in-situ density may not be known. The accuracy of this technique in this case
must be recognized when interpreting the results. However, for estimating the shear strength of
compacted backfill, more accurate results can be obtained, since the soil placement method, as well
as the in-situ density and moisture content, can be recreated in the laboratory with some degree of
confidence. The key in the latter case is the specimen size allowed by the testing device, as in many
cases, compacted fills have a significant percentage of gravel sized particles, requiring fairly large
test specimens (i.e., minimum 3 to 4 inch diameter, or narrowest dimension specimens of 3 to 4
inches).
Typically, a disturbed sample of the granular backfill material (or native material in the case of
obtaining supplementary information for back-analysis of existing slopes) is sieved to remove
particles that are too large for the testing device and test standard, and is compacted into a mold to
simulate the final density and moisture condition of the material. The specimens may or may not be
saturated after compacting them and placing them in the shear testing device, depending on the
condition that is to be simulated. In general, a drained test is conducted, or if it is saturated, the pore
pressure during shearing can be measured (possible for triaxial testing; generally not possible for
direct shear testing) to obtained drained shear strength parameters. Otherwise, the test is run slow
enough to be assured that the specimen is fully drained during shearing (note that estimating the
testing rate to assure drainage can be difficult). Multiple specimens tested using at least three
confining pressures should be tested to obtain a shear strength envelope. See Sabatini, et al. (2002)
for additional details.

5.6.1.2

O t h e r L a b o r a t o r y Test s

Tests to evaluate compressibility or permeability of existing subsurface deposits must be conducted


on undisturbed specimens, and sample disturbance must be kept to a minimum. See Sabatini, et al.
(2002) for additional requirements regarding these and other types of laboratory performance tests
that should be followed.

5.6.2

Correlations to Estimate Engineering Properties of


Soil

Correlations that relate in-situ index test results such as the SPT or CPT or laboratory soil index
testing may be used in lieu of, or in conjunction with, performance laboratory testing and backanalysis of site performance data to estimate input parameters for the design of the geotechnical
elements of a project. Since properties estimated from correlations tend to have greater variability
than measurement using laboratory performance data (see Phoon, et al., 1995), properties estimated
from correlation to in-situ field index testing or laboratory index testing should be based on multiple
measurements within each significant geologic unit (if the geologic unit is large enough to obtain
multiple measurements). A minimum of 3 to 5 measurements should be obtained from each geologic
unit as the basis for estimating design properties.
The drained friction angle of granular deposits should be determined based on the correlation
provided in Table 5 -12.
Table 5-12. Correlation of SPT N values to drained friction angle of granular soils (modified after
Bowles, 1977)

N1(60) from SPT


(blows/ft)

(deg)

Volume 1
Manual

ODOT Geotechnical Design


7

April 2010

<4
4
10
30
50

25-30
27-32
30-35
35-40
38-43

Experience should be used to select specific values within the ranges. In general, finer materials or
materials with significant silt-sized material will fall in the lower portion of the range. Coarser materials
with less then 5% fines will fall in the upper portion of the range.
Care should be exercised when using other correlations of SPT results to soil parameters. Some
published correlations are based on corrected values (N1(60)) and some are based on uncorrected
values (N). The designer should ascertain the basis of the correlation and use either N1(60) or N as
appropriate. Care should also be exercised when using SPT blow counts to estimate soil shear
strength if in soils with coarse gravel, cobbles, or boulders. Large gravels, cobbles, or boulders could
cause the SPT blow counts to be unrealistically high.
Correlations for other soil properties (other than as specifically addressed above for the soil friction
angle) as provided in Sabatini, et al. (2002) may be used if the correlation is well established and if
the accuracy of the correlation is considered regarding its influence if the estimate obtained from the
correlation in the selection of the property value used for design. Local geologic formation-specific
correlations may also be used if well established by data comparing the prediction from the
correlation to measured high quality laboratory performance data, or back-analysis from full scale
performance of geotechnical elements affected by the geologic formation in question.

Engineering Properties of Rock

Engineering properties of rock are generally controlled by the discontinuities within the rock mass and
not the properties of the intact material. Therefore, engineering properties for rock must account for
the properties of the intact pieces and for the properties of the rock mass as a whole, specifically
considering the discontinuities within the rock mass. A combination of laboratory testing of small
samples, empirical analysis, and field observations should be employed to determine the engineering
properties of rock masses, with greater emphasis placed on visual observations and quantitative
descriptions of the rock mass.
Rock properties can be divided into two categories: intact rock properties and rock mass properties.

Intact rock: Intact rock properties are determined from laboratory tests on small samples
typically obtained from coring, outcrops or exposures along existing cuts. Engineering
properties typically obtained from laboratory tests include specific gravity, unit weight,
ultrasonic velocity, compressive strength, tensile strength, and shear strength.

Rock mass properties: Rock mass properties are determined by visual examination and
measurement of discontinuities within the rock mass, and how these discontinuities will affect
the behavior of the rock mass when subjected to the proposed construction.

The methodology and related considerations provided by Sabatini, et al. (2002) should be used to
assess the design properties for the intact rock and the rock mass as a whole.

Volume 1
Manual

ODOT Geotechnical Design


8

April 2010

However, the portion of Sabatini, et al. (2002) that addresses the determination of fractured rock
mass shear strength parameters (Hoek and Brown, 1988) is outdated. The original work by Hoek
and Brown has been updated and is described in Hoek, et al. (2002).
The updated method uses a Geological Strength Index (GSI) to characterize the rock mass for the
purpose of estimating strength parameters, and has been developed based on re-examination of
hundreds of tunnel and slope stability analyses in which both the 1988 and 2002 criteria were used
and compared to field results. While the 1988 method has been more widely published in national
(e.g., FHWA) design manuals than has the updated approach provided in Hoek, et al. (2002),
considering that the original developers of the method have recognized the short-comings of the
1988 method and have reassessed it through comparison to actual rock slope stability data, the
Hoek, et al. (2002) is considered to be the most accurate methodology. Therefore the Hoek, et al.
(2002) method should be used for fractured rock mass shear strength determination. Note that this
method is only to be used for highly fractured rock masses in which the stability of the rock slope is
not structurally controlled.

Volume 1
Manual

ODOT Geotechnical Design


9

April 2010


5.8.1

Final Selection of Design Values


Overview

After the field and laboratory testing is completed, the geotechnical designer should review the quality
and consistency of the data, and should determine if the results are consistent with expectations.
Once the lab and field data have been collected, the process of final material property selection
begins. At this stage, the geotechnical designer generally has several sources of data consisting of
that obtained in the field, laboratory test results and correlations from index testing. In addition, the
geotechnical designer may have experience based on other projects in the area or in similar soil/rock
conditions. Therefore, if the results are not consistent with each other or previous experience, the
reasons for the differences should be evaluated, poor data eliminated and trends in data identified. At
this stage it may be necessary to conduct additional performance tests to try to resolve
discrepancies.

Geotechnical Design Property Assessment


As stated in Section 5.1, the focus of geotechnical design property assessment and final selection is
on the individual geologic strata identified at the project site. A geologic stratum is characterized as
having the same geologic depositional history and stress history, and generally has similarities
throughout the stratum in its density, source material, stress history, and hydrogeology. All of the
information that has been obtained up to this point including preliminary office and field
reconnaissance, boring logs, CPT soundings etc., and laboratory data are used to determine soil and
rock engineering properties of interest and develop a subsurface model of the site to be used for
design. Data from different sources of field and lab tests, from site geological characterization of the
site subsurface conditions, from visual observations obtained from the site reconnaissance, and from
historical experience with the subsurface conditions at or near the site must be combined to
determine the engineering properties for the various geologic units encountered throughout the site.
However, soil and rock properties for design should not be averaged across multiple strata, since the
focus of this property characterization is on the individual geologic stratum. Often, results from a
single test (e.g. SPT N-values) may show significant scatter across a site for a given soil/rock unit.
Data obtained from a particular soil unit for a specific property from two different tests (e.g. field vane
shear tests and lab UU tests) may not agree. Techniques should be employed to determine the
validity and reliability of the data and its usefulness in selecting final design parameters. After a
review of data reliability, a review of the variability of the selected parameters should be carried out.
Variability can manifest itself in two ways: 1) the inherent in-situ variability of a particular parameter
due to the variability of the soil unit itself, and 2) the variability associated with estimating the
parameter from the various testing methods. From this step, final selection of design parameters can
commence, and from there completion of the subsurface profile.

5.8.2

Data Reliability and Variability

Inconsistencies in data should be examined to determine possible causes and assess any mitigation
procedures that may be warranted to correct, exclude, or downplay the significance of any suspect
data. Chapter 8 of Sabatini, et al. (2002) outlines step-by-step procedures for analyzing data and
resolving inconsistencies.

Volume 1
Manual

ODOT Geotechnical Design


10

April 2010

5.8.3

Final Property Selection

The final step is to incorporate the results of the previous section into the selection of values for
required design properties. Recognizing the degree of variability discussed in the previous section,
the potential impact of that variability (or uncertainty) on the level of safety in the design, and on
potential cost and constructability impacts, should be assessed. If the impact of this uncertainty is
likely to be significant, parametric analyses should be conducted, or more data could be obtained to
help reduce the uncertainty. Since the sources of data that could be considered may include
measured laboratory data, field test data, performance data, and other previous experience with the
geologic unit(s) in question, it will not be possible to statistically combine all this data together to
determine the most likely property value.
Engineering judgment, combined with parametric analyses as needed, will be needed to make the
final assessment and determination of each design property. This assessment should include a
decision as to whether the final design value selected should reflect the interpreted average value for
the property, or a value that is somewhere between the most likely average value and the most
conservative estimate of the property. Design property selection should achieve a balance between
the desire for design safety and the cost effectiveness and constructability of the design. In some
cases, the selection of conservative design properties could result in very conservative designs that
are un-constructible (e.g., using very conservative design parameters resulting in a pile foundation
that must be driven deep into a very dense soil unit that in reality is too dense to penetrate with
available equipment).
Note that in Chapter 8, where reliability theory was used to establish load and resistance factors, the
factors were developed assuming that mean values for the design properties are used. However,
even in those cases, design values that are more conservative than the mean may still be
appropriate, especially if there is an unusual amount of uncertainty in the assessment of the design
properties due, for example to highly variable site conditions, lack of high quality data to assess
property values, or due to widely divergent property values from the different methods used to assess
properties within a given geologic unit.
Depending on the availability of soil or rock property data and the variability of the geologic strata
under consideration, it may not be possible to reliably estimate the average value of the properties
needed for design. In such cases, the geotechnical designer may have no choice but to use a more
conservative selection of design parameters to mitigate the additional risks created by potential
variability or the paucity of relevant data. Note that for those resistance factors that were determined
based on calibration by fitting to allowable stress design, this property selection issue is not relevant,
and property selection should be based on the considerations discussed previously.
The process and examples to make the final determination of properties to be used for design
provided by Sabatini, et al. (2002) should be followed.

5.8.4

Development of the Subsurface Profile

While Section 5.8 generally follows a sequential order, it is important to understand that the selection
of design values and production of a subsurface profile is more of an iterative process. The
development of design property values should begin and end with the development of the subsurface
profile. Test results and boring logs will likely be revisited several times as the data is developed and
analyzed before the relation of the subsurface units to each other and their engineering properties
are finalized.
Volume 1
Manual

ODOT Geotechnical Design


11

April 2010

The ultimate goal of a subsurface investigation is to develop a working model that depicts major
subsurface layers exhibiting distinct engineering characteristics.
The end product is the subsurface profile, a two dimensional depiction of the site stratigraphy. The
following steps outline the creation of the subsurface profile:
1. Complete the field and lab work and incorporate the data into the preliminary logs.
2. Lay out the logs relative to their respective field locations and compare and match up the
different soil and rock units at adjacent boring locations, if possible. However, caution should
be exercised when attempting to connect units in adjacent borings, as the geologic
stratigraphy does not always fit into nice neat layers. Field descriptions and engineering
properties will aid in the comparisons.
3. Group the subsurface units based on engineering properties.
4. Create cross sections by plotting borings at their respective elevations and positions
horizontal to one another with appropriate scales. If appropriate, two cross sections should be
developed that are at right angles to each other so that lateral trends in stratigraphy can be
evaluated when a site contains both lateral and transverse extents (i.e. a building or large
embankment).
5. Analyze the profile to see how it compares with expected results and knowledge of geologic
(depositional) history. Have anomalies and unexpected results encountered during
exploration and testing been adequately addressed during the process? Make sure that all of
the subsurface features and properties pertinent to design have been addressed.

5.8.5

Selection of Design Properties for Engineered


Materials

This section provides guidelines for the selection of properties that are commonly used on ODOT
projects such as engineered fills. The engineering properties are based primarily on gradation and
compaction requirements, with consideration of the geologic source of the fill material typical for the
specific project location. For materials such as common borrow where the gradation specification is
fairly broad, a wider range of properties will need to be considered.

5.8.5.1

Borrow Material

The standard specification for Borrow Material, section 00330.12, states it may be virtually any soil or
aggregate either naturally occurring or processed which is free of unsuitable materials. Follow these
guidelines:

On ODOT projects, Borrow Material which meets the criteria for Moisture-Density Testable
Material is compacted to at least 95 percent of maximum density based on the Standard
Proctor in accordance with 00330.43(b) (2-b). Borrow Material which is a Non-Moisture
Density Testable Material is typically compacted in accordance with the procedure described
in 00330.43(c).

Because of the variability of the materials that may be used as Borrow Material, the
estimation of an internal friction angle and unit weight should be based on the actual material
used.

Volume 1
Manual

ODOT Geotechnical Design


12

April 2010

For non-plastic materials, the friction angle may be in the 30 to 34 degree range, and the unit
weight may be in the 115 to 130 pcf range.

Lower range values should be used for finer grained materials compacted to 90 percent of
maximum density.

In general during design, the specific source of borrow is not known. Therefore, it is not prudent to
select a design friction angle that is near or above the upper end of the range unless the geotechnical
designer has specific knowledge of the source(s) likely to be used, or unless quality assurance shear
strength testing is conducted during construction. Borrow material will likely have a high enough fines
content to be moderately to highly moisture sensitive. This moisture sensitivity may affect the design
property selection if it is likely that placement conditions are likely to be marginal due to the timing of
construction.

5.8.5.2

S e l e c t G r a n u l a r B a c kf i l l

The standard specification for Select Granular Backfill, section 00330.14, ensures that the mixture
will be granular and contain at least a minimal amount of gravel size material. The materials are likely
to be poorly graded sand and contain enough fines to be moderately moisture sensitive. The
following applies:

Select Granular Backfill is not an all-weather material. Select Granular Backfill gradation
indicates that drained friction angles of 34 to 38 degrees are possible when the soil is well
compacted.

Relatively clean sands in a loose state will likely have drained friction angles of 30 to 35
degrees. Unit weights will be in the 120 to 130 pcf range for all the Select Granular Backfill
materials. However, these values are highly dependent on the geologic source of the
material. Windblown, beach, or alluvial sands that have been rounded through significant
transport could have significantly lower shear strength values.

Reject and scalped materials from processing could also have relative low friction angles
depending on the uniformity of the material and the degree of rounding in the soil particles.

In general, during design, the specific source of borrow is not known. Therefore, it is not prudent to
select a design friction angle that is near or above the upper end of the range unless the geotechnical
designer has specific knowledge of the source(s) likely to be used or unless quality assurance shear
strength testing is conducted during construction. Select Granular Backfill with significant fines
content may sometimes be modeled as having a temporary or apparent cohesion value from 50 to
200 psf. If a cohesion value is used, the friction angle should be reduced so as not to increase the
overall strength of the material. For long term analysis, all the Granular Materials should be modeled
with no cohesive strength.

5.8.5.3

Select Stone Backfill

The standard specification for Select Stone Backfill, section 00330.15, should ensure reasonably well
graded sand and gravel. Maximum fines content is not specified, so the material may be moisture
sensitive. In very wet conditions, material with lower fines content should be used. The Select Stone
Backfill specification indicates that internal angles of friction up to 40 degrees are possible, and that
shear strength values less than 36 degrees are not likely. However, lower shear strength values are
possible for Select Stone Backfill from naturally occurring materials obtained from non-glacially
derived sources such as wind blown or alluvial deposits. In many cases, processed materials are
Volume 1
Manual

ODOT Geotechnical Design


13

April 2010

used for Select Stone Backfill, and in general, this processed material has been crushed, resulting in
rather angular particles and high soil friction angles. Unit weights of 130 to 140 pcf are possible if very
well graded. In general, during design, the specific source of borrow is not known. Therefore, it is not
prudent to select a design friction angle that is near or above the upper end of the range unless the
geotechnical designer has specific knowledge of the source(s) likely to be used or unless quality
assurance shear strength testing is conducted during construction.

5.8.5.4

S t o n e E m b a n k m en t M at e r i a l

Stone Embankment Material, standard specification section 00330.16, is considered an all-weather


material. Compactive effort is based on a method specification. Because of the nature of the material,
compaction testing is generally not feasible. The specification allows for a broad range of material
and properties such that the internal friction angle and unit weight can vary considerably based on
the amount and type of rock in the fill. For compacted rock embankments constructed with Stone
Embankment Material:

Internal friction angles of up to 45 degrees may be reasonable.

Unit weights for rock embankments generally range from 130 to 140 pcf.

Durability is major issue with this material. Rock excavated from cuts consisting of siltstone,
sandstone and claystone may break down during the compaction process, resulting in less coarse
material. Also, if the rock is weak, failure may occur through the rock fragments rather than around
them. In these types of materials, the strength parameters may resemble those of embankments
constructed from Borrow Materials. For existing embankments, the soft rock may continue to weather
with time, if the embankment materials continue to become wet. Inadequate slope stability and
excessive settlement of embankments with non-durable materials are the long term effects of using
weak rock materials without proper placement and compaction.

5.8.5.5

Woo d F i b e r

Wood fiber fills have been used by ODOT for fill heights up to about 20 feet. The wood fiber has
generally been used as lightweight fill material in emergency repair situations because wet weather
does not affect the placement and compaction of the embankment. Only fresh wood fiber should be
used to prolong the life of the fill, and the maximum particle size should be 6 inches or less. The
wood fiber is generally compacted in lifts of about 12 inches with two or more passes of a track dozer.
Presumptive design values of 50 pcf for unit weight and an internal angle of friction of about 40
degrees may be used for the design of the wood fiber fills (Allen et al., 1993).
To mitigate the effects of leachate, the amount of water entering the wood should be minimized.
Generally, topsoil caps of about 2 feet in thickness are used. The pavement section should be a
minimum of 2 feet (a thicker section may be needed depending on the depth of wood fiber fill). Wood
fiber fill will experience creep settlement for several years and some pavement distress should be
expected during that period. Additional information on the properties and durability of wood fiber fill is
provided in Kilian and Ferry (1993).

5.8.5.6

Geofoam

Geofoam has not been used as lightweight fill on ODOT projects, but there may be projects that will
incorporate it in the future. In contrast, WSDOT has had about 10 years of experience with Geofoam
in embankment construction. Geofoam ranges in unit weight from about 1 to 2 pcf. The Geofoam
material is made from expanded polystyrene (EPS) and is manufactured according to ASTM
Volume 1
Manual

ODOT Geotechnical Design


14

April 2010

standards for minimum density (ASTM C 303), compressive strength (ASTM D 1621) and water
absorption (ASTM C 272). Type I and II Geofoam are generally used in highway applications. Bales
of recycled industrial polystyrene waste are also available. These bales have been used to construct
temporary haul roads over soft soil. However, these bales should not be used in permanent
applications.

Volume 1
Manual

ODOT Geotechnical Design


15

April 2010

References

Allen, T. M., Kilian, A. P., 1993, Use of Wood Fiber and Geotextile Reinforcement to Build
Embankment Across Soft Ground, Transportation Research Board Record 1422.
Hoek, E., and Brown, E.T. 1988. The Hoek-Brown Failure Criterion a 1988 Update. Proceedings,
15th Canadian Rock Mechanics Symposium, Toronto, Canada.
Hoek, E., Carranza-Torres, C., and Corkum, B., 2002, Hoek-Brown Criterion 2002 Edition,
Proceedings NARMS-TAC Conference, Toronto, 2002, 1, pp. 267-273.
Kilian, A. P., Ferry, C. D., 1993, Long Term Performance of Wood Fiber Fills, Transportation Research
Board Record 1422.
Mayne, P. W., Christopher, B. R., DeJong, J., 1997, FHWA-HI-97-021, Subsurface Investigations,
NHI course manual #13201.
Phoon, K.-K., Kulhawy, F. H., Grigoriu, M. D., 1995, Reliability-Based Design of Foundations for
Transmission Line Structures, Report TR-105000, Electric Power Research Institute, Palo Alto, CA.
Sabatini, P. J., Bachus, R. C., Mayne, P. W., Schneider, T. E., Zettler, T. E., U.S. Department of
Transportation, Federal Highway Administration Evaluation of Soil and Rock Properties,
Geotechnical Engineering Circular No. 5, FHWA-IF-02-034, April, 2002.

Volume 1
Manual

ODOT Geotechnical Design


16

April 2010

Chapter

Seismic Design

6.1. General
This chapter describes ODOTs standards and policies regarding the geotechnical aspects of the
seismic design of ODOT projects. The purpose is to provide geotechnical engineers and engineering
geologists with specific seismic design guidance and recommendations not found in other standard
design documents used for ODOT projects. Complete design procedures (equations, charts, graphs,
etc.) are usually not provided unless necessary to supply, or supplement, specific design information,
or if they are different from standards described in other references. This chapter also describes what
seismic recommendations should typically be provided by the geotechnical engineer in the
Geotechnical Report.

6.1.1 Seismic Design Standards


The seismic design of ODOT bridges shall follow methods described in the most currently adopted
edition of the AASHTO Guide Specifications for LRFD Seismic Bridge Design (AASHTO, 2009)
supplemented by the Bridge Design and Drafting Manual (BDDM) and the recommendations
supplied in this chapter.. Refer to the ODOT BDDM for additional design criteria and guidance
regarding the use of the new Guide Specification on bridge projects. The term AASHTO as used in
this chapter refers to AASHTO LRFD design methodology. For seismic design of new buildings the
2003 International Building Code (IBC) (International Code Council, 2002) should be used. In
addition to these standards, the manuals listed below may be referenced for additional design
guidance in seismic design for issues and areas not addressed in detail in the AASHTO
specifications or this chapter. Unless otherwise noted, the standards and policies described in this
chapter supersede those described in the referenced documents.

Design Guidance: Geotechnical Earthquake Engineering For Highways, Geotechnical


Engineering Circular No. 3, Volumes I & II, FHWA-SA-97-076&077.

This FHWA document provides design guidance for geotechnical earthquake engineering for
highways. Specifically, this document provides guidance on earthquake engineering fundamentals,
seismic hazard analysis, ground motion characterization, site characterization, seismic site response
analysis, seismic slope stability, liquefaction, and seismic design of foundations and retaining walls.

Volume 1
17

ODOT Geotechnical Design Manual


April 2010

The document also includes design examples (Volume II) for typical geotechnical earthquake
engineering analyses.
For liquefaction analysis, embankment deformation estimates and bridge damage assessment the
following two documents should be referenced:

Assessment and Mitigation of Liquefaction Hazards to Bridge Approach Embankments in


Oregon, Dickenson, S., et al., Oregon State University, Department of Civil, Construction
and Environmental Engineering, SPR Project 361, November, 2002.

Recommended Guidelines For Liquefaction Evaluations Using Ground Motions From


Probabilistic Seismic Hazard Analysis, Dickenson, S., Oregon State University, Department
of Civil, Construction and Environmental Engineering, Report to ODOT, June, 2005.

The above two documents are available on the ODOT Geo-Environmental web page. In light of the
continuous advances being made in evaluating the impact of liquefaction hazards and ground
failures on structures the above two references should be supplemented with the most up to date
technical information and guidelines.

NCHRP Report 472: The National Cooperative Highway Research Program Report 472
(2002), Comprehensive Specifications for the Seismic Design of Bridges, is a report
containing the findings of a study completed to develop recommended specifications for
seismic design of highway bridges. The report covers topics including design earthquakes
and performance objectives, foundation design, liquefaction hazard assessment and design,
and seismic hazard representation.

United States Geological Survey (USGS) Website.


The USGS National Seismic Hazard Maps website is a valuable tool for characterizing the
seismic hazard for a specific site. This site provides the results of Probabilistic Seismic
Hazard Analyses (PSHA) in the form of the Uniform Seismic Hazard, which reflects the
contribution of all seismic sources in the region on the ground motion parameters. The
website provides ground motion parameters (Peak Ground Acceleration (PGA), and
acceleration response spectral ordinates between 0.1 and 5.0 seconds) on Site Class B rock
for various return periods, specified as a percentage probability of exceedance in a given
exposure interval, in years. The website provides PGA and spectral acceleration ordinates at
periods of 0.2 and 1.0 second for risk levels of 5 and 10 percent probabilities of exceedance
(PE) in 50 years for use in the development of response spectra, in accordance with
AASHTO. This risk level of 5 and 10 percent PE corresponds to approximately 7 and 14
percent PE in 75 years, respectively. The website also provides interactive deaggregation of
a sites probabilistic seismic hazard. The deaggregation is useful for demonstrating the
relative contribution of regional seismic sources, in terms of magnitude and source-to-site
distance, on the seismic hazard at a site. De-aggregation is particularly useful for
demonstrating how the ground motion parameters generated by individual sources compare
to the mean motions determined for the Uniform Seismic Hazard.

WSDOT Geotechnical Design Manual, M46-03.01, November, 2008


o
http://www.wsdot.wa.gov/publications/manuals/fulltext/M4603/Geotech.pdf

6.1.1

Background

In light of the complexity of seismic foundation design, continuous enhancements to analytical and
empirical methods of evaluation are being made as more field performance data is collected and
Volume 1
18

ODOT Geotechnical Design Manual


April 2010

research advances the state of knowledge. New methods of analysis and design are continuously
being developed and therefore it is considered prudent to not be overly prescriptive in defining
specific design methods for use in the seismic design process. However, a standard of practice
needs to be established within the geotechnical community regarding minimum required design
criteria for seismic design. It is well recognized that these standards are subject to change in the
future as a result of further research and studies. This chapter will be continually updated as more
information is obtained, new design codes are approved and better design methods become
available.
Significant engineering judgment is required throughout the entire seismic design process. The
recommendations provided herein assume the geotechnical designer has a sound education and
background in basic earthquake engineering principles. These recommendations are not intended to
be construed as complete or absolute. Each project is different in some way and requires important
decisions and judgments be made at key stages throughout the design process. The applicability of
these recommended procedures should be continually evaluated throughout the design process.
Peer review may be required to assist the design team in various aspects of the seismic hazard and
earthquake-resistant design process.
Earthquakes often result in the transfer of large axial and lateral loads from the bridge superstructure
into the foundations. At the same time, foundation soils may liquefy, resulting in a loss of soil strength
and foundation capacity. Under this extreme event condition it is common practice to allow the
foundations to be loaded up to the nominal (ultimate) foundation resistances (allowing resistance
factors as high as 1.0). This design practice requires an increased emphasis on quality control during
the construction of bridge foundations since we are now often relying on the full, unfactored nominal
resistance of each foundation element to support the bridge during the design seismic event.
In addition to seismic foundation analysis, seismic structural design also involves an analysis of the
soil-structure interaction between foundation materials and foundation structure elements. Soilstructure interaction is typically performed in bridge design by modeling the foundation elements
using equivalent linear springs. Some of the recommendations presented herein relate to bridge
foundation modeling requirements and the geotechnical information the structural designer needs in
order to do this analysis. Refer to Section 1.1.4 of the ODOT Bridge Design and Drafting Manual
(BDDM) for more information on bridge foundation modeling procedures.

6.1.2

Responsibility of the Geotechnical Designer

The geotechnical designer is responsible for providing geotechnical/seismic input parameters to the
structural engineers for their use in design of the transportation infrastructure. Specific elements to be
addressed by the geotechnical designer include the design ground motion parameters, site
response, geotechnical design parameters and geologic hazards. The geotechnical designer is also
responsible for providing input for evaluation of soil-structure interaction (foundation response to
seismic loading), earthquake induced earth pressures on retaining walls, and an assessment of the
impacts of geologic hazards on the structures. Refer to Chapter 21 for geotechnical seismic design
reporting requirements.
The seismic geologic hazards to be evaluated include fault rupture, liquefaction, ground failure
including flow slides and lateral spreading, ground settlement, and instability of natural slopes and
earth structures. The seismic performance of tunnels is a specialized area of geotechnical
earthquake engineering not specifically addressed in this guidance document; however the ground
motion parameters determined in the seismic hazard analyses outlined herein may form the basis for
tunnel stability analyses (e.g., rock fall adjacent to portals and in unlined tunnels, performance of
tunnel lining). The risk associated with seismic geologic hazards shall be evaluated by the
geotechnical designer following the methods described in this chapter.
Volume 1
19

ODOT Geotechnical Design Manual


April 2010

Seismic
Design
Requirements

6.2.1

New Bridges

Performance

A two-level approach is used in ODOT for the seismic design of all new bridges. The seismic design
of ODOT bridges is evaluated in terms of performance requirements for ground motions having
average return periods of approximately 500-years and 1000-years. The 500-year and 1000-year
return period ground motions have probabilities of exceedance of approximately 14% and 7% in 75
years respectively. For a 50 year time period, the probabilities of exceedance of the 500 and 1000
year ground motions are approximately 10% and 5% respectively. The seismic foundation design
requirements, including approach embankments, shall be consistent with meeting the current ODOT
Bridge Engineering Section seismic design criteria. Excerpts from those criteria are summarized as
follows from the BDDM:
1000-year No-Collapse Criteria: Design all bridges for 1000-year return period ground
motions (7% probability of exceedance in 75 years) under No Collapse criteria.
Under this level of shaking, the bridge and approach structures, bridge foundation and
approach fills must be able to withstand the forces and displacements without collapse of any
portion of the structure. In general, bridges that are properly designed and detailed for
seismic loads can accommodate relatively large deflections without the danger of collapse. If
large embankment displacements (lateral spread) or overall slope failure of the end fills are
predicted, the impacts on the bridge end bent, abutment walls and interior piers should be
evaluated to see if the impacts could potentially result in collapse of any part of the structure.
Slopes adjacent to a bridge or tunnel should be evaluated if their failure could result in
collapse of a portion or all of the structure.
500-year Serviceability Criteria: In addition to the 1000-year No Collapse criteria,
design all bridges to remain Serviceable after subjected to 500-year return period ground
motions (14% probability of exceedance in 75 years).
Under this level of shaking, the bridge and approach fills, are designed to remain in service
shortly after the event (after the bridge has been properly inspected) to provide access for
emergency vehicles. In order to do so, the bridge is designed to respond semi-elastically
under seismic loads with minimal damage. Some structural damage is anticipated but the
damage should be repairable and the bridge should be able to carry emergency vehicles
immediately following the earthquake. This holds true for the approach fills leading up to the
bridge.
Approach fill settlement and lateral displacements should be minimal to provide for immediate
emergency vehicle access for at least one travel lane. For mitigation purposes approach fills
are defined as shown on Figure 6 -13. As a general rule of thumb, an estimated lateral
embankment displacement of up to 1 foot is considered acceptable in many cases as long as
the serviceable performance criteria described above can be met. Vertical settlements on
the order of 6 to 12 may be acceptable depending on the roadway geometry and
anticipated performance of the bridge end panels. Bridge end panels are required on all state
highway bridge projects (per BDDM) and should be evaluated for their ability to withstand the
anticipated embankment displacements and settlement and still provide the required level of
performance. These displacement criteria are to serve as general guidelines only and
engineering judgment is required to determine the final amounts of acceptable displacement
that will meet the desired criteria. It should be noted that these estimated displacements are
Volume 1
20

ODOT Geotechnical Design Manual


April 2010

not at all precise values and may easily vary by factors of 2 to 3 depending on the analysis
method(s) used. The amounts of allowable vertical and horizontal displacements should be
decided on a case-by-case basis, based on discussions and consensus between the bridge
designer and the geotechnical designer and perhaps other project personnel.
In addition to bridge and approach fill performance, embankments through which cut-and-cover
tunnels are constructed should be designed to remain stable during the design seismic event
because of the potential for damage or possible collapse of the structure should they fail.
Approach embankments and structure foundations should be designed to meet the above
performance requirements. Unstable slopes such as active or potential landslides, and other seismic
hazards such as liquefaction, lateral spread, post-earthquake settlement and downdrag may require
mitigation measures to ensure that the structure meets these performance requirements. Refer to
Chapter 11 for guidance on approved ground improvement techniques to use in mitigating these
hazards.

6.2.2

Bridge Widenings

For the case where an existing bridge is to be widened, the foundations for the widened portion of the
bridge and bridge approaches should be designed to remain stable and meet the same bridge
performance criteria as new bridges. In addition, if the existing bridge foundation is not stable and
could cause collapse of the bridge widening, or if liquefiable soils are present, to the extent practical,
measures should be taken to prevent collapse of the existing bridge during the design seismic event.
If foundation retrofit or liquefaction mitigation is necessary to meet the performance criteria, these
designs shall be reviewed and approved by the HQ Bridge Section. The foundations for the widening
should be designed in a way that the seismic response of the bridge widening can be made
compatible with the seismic response of the existing bridge as stabilized in terms of foundation
deformation and stiffness. If it is not feasible to stabilize the existing bridge such that it will not cause
collapse of the bridge widening during the design seismic event, consideration should be given to
replacing the existing bridge rather than widening it.

6.2.3

Bridge Abutments and Retaining Walls

Seismic design performance objectives for retaining walls depend on the function of the retaining wall
and the potential consequences of failure.
There are four retaining wall categories, as defined in Section 15.2.1. The seismic design
performance objectives for these four categories are listed below:

Bridge Abutments: Bridge Abutments are considered to be part of the bridge, and
shall meet the seismic design performance objectives for the bridge see
Section 6.2.1.

Bridge Retaining Walls: Design all Bridge Retaining Walls for 1000-year return
period ground motions under the No Collapse bridge criteria. Under this level of
shaking, the Bridge Retaining Wall must be able to withstand seismic forces and
displacements without failure of any part of the wall or collapse of any part of the
bridge which it supports. Bridge Retaining Walls shall be designed for overall stability
under these seismic loading conditions, including anticipated displacements
associated with liquefaction. Mitigation to achieve overall stability may be required.
In addition, design all Bridge Retaining Walls for 500-year return period ground motions
under the Serviceability bridge criteria. Under this level of shaking, Bridge Retaining
Wall movement must not result in unacceptable performance of the bridge or bridge

Volume 1
21

ODOT Geotechnical Design Manual


April 2010

approach fill, as described under the 500-Year Serviceability criteria in


Section 6.2.1.

Highway Retaining Walls: Design all Highway Retaining Walls for 1000-year return
period ground motions. Under this level of shaking, the Highway Retaining Wall must
be able to withstand seismic forces and displacements without failure of any part of
the Highway Retaining Wall. Highway Retaining Walls shall be designed for overall
stability under these seismic loading conditions, including anticipated displacements
associated with liquefaction. Mitigation to achieve overall stability may be required

Minor Retaining Walls: Minor Retaining Wall systems have no seismic design
requirements.

The policy to design all Highway Retaining Walls to meet overall stability requirements for seismic
design may not be practical at all wall locations. Where it is not practical to design a Highway
Retaining Wall for overall stability under seismic loading, and where a failure of this type would not
endanger the public, impede emergency and response vehicles along essential lifelines, or have an
adverse impact on another structure, the local Region Tech Center will evaluate practicable
alternatives for improving the seismic resistance and performance of the retaining wall.
In general, retaining walls and bridge abutments should not be built on or near landslides or other
areas that are marginally stable under static conditions. However, if site conditions and project
constraints provide no cost effective or technical alternative, the local Region Tech Center will
evaluate, on a case-by-case basis, the possible placement of these structures in these locations, as
well as requirements for global (overall) instability of the landslide during the design seismic event.

6.2.4

Embankments and Cut Slopes

Cut slopes, fill slopes, and embankments are generally not evaluated for seismic instability unless
they directly affect a bridge, highway retaining wall or other structure. Bridge approach fills should
always be evaluated for stability and settlement, especially if they are relied upon to provide passive
soil resistance behind the abutment (Earthquake-Resisting System). Seismic instability associated
with routine cuts and fills are typically not mitigated due to the high cost of applying such a design
policy uniformly to all slopes statewide. If failure and displacement of existing slopes, embankments
or cut slopes, due to seismic loading, could adversely impact an adjacent structure, these areas
should be considered for stabilization. Such impacts should be evaluated in terms of meeting the
performance criteria described in Section 6.2.

Ground Motion Parameters

The ground motion parameters to be used in design are currently based on the 2002 USGS National
Seismic Hazard maps for the Pacific Northwest region. The USGS seismic hazard mapping project
provides the results of probabilistic seismic hazard analysis (PSHA) at the regional scale. The ground
motion maps for the 500 and 1000-year return periods are available in the ODOT Bridge Design and
Drafting Manual (BDDM).
Ground motion parameters for the 2002 USGS hazard maps are also available on the USGS
website at:
http://earthquake.usgs.gov/hazmaps/
The designer should review the basis of these hazard maps and have a thorough understanding of
the data they represent and the methods used for their development. The 2008 USGS National
Volume 1
22

ODOT Geotechnical Design Manual


April 2010

Seismic Hazard maps are currently under review by the ODOT Bridge Section and are not approved
for use on ODOT projects at this time.
The BDDM maps provide Peak Ground Acceleration (PGA), 0.20 sec. and 1.0 sec. spectral
accelerations scaled in contour intervals of 0.01g. Ground motion values can be obtained from the
USGS website by selecting the Interactive Deaggregations link under the Seismic Hazard
Analysis Tools heading. The PGA and spectral accelerations can then be obtained by entering the
latitude and longitude of the site and the desired probability of exceedance (i.e., 5% in 50 years for
the 1000 year return event). It should be noted that the PGA obtained from these maps is actually the
Peak Bedrock Acceleration (i.e., Site Class B), and does not include, or take into account, any local
soil amplification effects. See Section 6.5.1 for the development of design ground motion data.

6.3.1

Site Specific Probabilistic Seismic Hazard Analysis

Ground motion parameters are also sometimes determined from a site specific Probabilistic Seismic
Hazard Analysis (PSHA). A site specific probabilistic hazard analysis may be considered when:

New information about one or more active seismic sources that affect the site has
become available since the most recent USGS/AASHTO Seismic Hazard Maps were
developed (2006), and the new seismic source information will result in a significant
change to the seismic hazard at the site. The existence of the Next Generation
Attenuation (NGA) relationships shall not be the sole justification for conducting a
PSHA.

The site is located within 6 miles of a known active fault capable of producing at least
a magnitude 5 earthquake. For these cases, near-fault ground motion effects
(directivity, directionality) were not explicitly modeled in the development of national
ground motion maps, and the code/specification based hazard level may be
significantly unconservative. These near-fault effects are normally only considered
for essential or critical structures.

The size or importance of the bridge is such that a lower probability of exceedance
(and therefore a longer return period) should be considered.

It should be noted that the site-specific PSHA is often uncoupled from subsequent site-specific
ground response analyses, which are described in Section 6.5.1.3. A site specific probabilistic
hazard analysis focuses on the spatial and temporal occurrence of earthquakes, and evaluates all of
the possible earthquake sources contributing to the seismic hazard at a site with the purpose of
developing ground motion data consistent with a specified uniform hazard level. The analysis takes
into account all seismic sources that may affect the site and quantifies the uncertainties associated
with the seismic hazard, including the location of the source, extent and geometry, maximum
earthquake magnitudes, rate of seismicity, and estimated ground-motion parameters. The result of
the analysis is a uniform hazard acceleration response spectrum that is based on a specified uniform
hazard level or probability of exceedance within a specified time period (i.e., 7% probability of
exceedance in 75 years). The PSHA is routinely performed to yield ground motion parameters for
bedrock (Site Class B) sites. The influence of the soil deposits at the site on the ground motion
characteristics is subsequently evaluated using the results of the PSHA for bedrock conditions. A
ground response analysis commonly utilizes computer programs such as SHAKE or D-MOD to
model the response of a soil column subjected to bedrock ground motion. The input bedrock motion
is in the form of time-histories that are typically matched (or scaled) to the bedrock response
spectrum developed from the probabilistic hazard analysis.
A site specific probabilistic hazard analysis is typically not performed on routine ODOT projects. If
such an analysis is desired for the design of ODOT bridge projects the HQ Bridge Engineering
Volume 1
23

ODOT Geotechnical Design Manual


April 2010

Section must approve the justification and procedures for conducting the analysis and the analysis
must be reviewed by an independent source approved by the HQ Bridge Engineering Section. The
review and approval of the PSHA will be coordinated with the region geotechnical engineer.

6.3.2

Magnitude and PGA for Liquefaction Analysis

Earthquake engineering evaluations that address repeated (cyclic) loading and failure of soils must
include estimates of the intensity and duration of the earthquake motions. In soils, liquefaction and
cyclic degradation of soil stiffness/strength represent fatigue failures that often impact bridge
structures. In practice-oriented liquefaction analysis, the intensity of the cyclic loading is related to the
PGA and/or cyclic stress ratio, and the duration of the motions is correlated to the magnitude of the
causative event. The PGA and magnitude values selected for the analysis should represent realistic
ground motions associated with specific, credible scenario earthquakes. The PGA values obtained
from the USGS web site represent the mean values of all of the sources contributing to the hazard
at the site for a particular recurrence interval. These mean PGA values should not typically be used
for liquefaction analysis unless the ground motions at the site are dominated by a single source, as
demonstrated in the PSHA deaggregation. Otherwise, the mean PGA values may not represent
realistic ground motions resulting from known sources affecting the site. Additionally, the mean
magnitude provided by PSHA should not be used as the causative event as this often averages the
magnitude of large Cascadia Subduction Zone earthquakes and the magnitude of the smaller, local
crustal events with a resulting magnitude that is not representative of any seismic source in the
region. For this reason the modal event(s), designated as Magnitude and Distance (M-R) pairs,
should be evaluated individually.

6.3.3

Deaggregation of Seismic Hazard

A deaggregation of the total seismic hazard should be performed to find the principal individual
sources contributing to the seismic hazard at the site. As a general rule of thumb, all sources that
contribute more than about 5% to the hazard should be evaluated. However, sources that contribute
less than 5% may also be sources to consider since they may still significantly affect the liquefaction
analysis or influence portions of the sites response spectra. The relative contribution of all considered
sources, in terms of magnitude and distance, on PGA and on spectral accelerations at 9 different
frequencies (or periods) of structural vibration can be readily evaluated using the results of the USGS
seismic hazard mapping tools and deaggregation capabilities available on-line.
It is recommended that the relative contributions of all of the following sources be considered when
performing liquefaction and ground deformation hazards:
(1) Cascadia Subduction Zone mega-thrust earthquakes
(2) Deep, intraslab Benioff Zone earthquakes such as the 1949 and 1965 Puget Sound,
and 2001 Nisqually earthquakes
(3) Shallow crustal earthquakes associated with mapped faults
(4) Regional background seismicity and randomly occurring earthquakes that are not
associated with mapped faults (gridded seismicity)
Deaggregation of the seismic hazard will provide the Magnitude (M) and Distance (R) of each source
contributing to the hazard at the site. These M & R values can then be utilized with ground motion
attenuation relationships to obtain bedrock PGA values at the site due to the individual sources. It is
recommended that more than one attenuation relationship be used to estimate ground motion
parameters for each of the primary seismic sources in Oregon (i.e., Cascadia Subduction Zone
events, and shallow crustal events). The use of three to four attenuation relationships is common in
Volume 1
24

ODOT Geotechnical Design Manual


April 2010

practice. In order to facilitate direct comparison of the ground motion parameters with the USGS
seismic hazard mapping results it is necessary to employ the same attenuation relationships that
were used in developing the 2002 USGS seismic hazard maps. These attenuation relationships are
summarized below. Additional attenuation relationships, appropriate for the style of faulting, can be
used at the discretion of the geotechnical engineer.
Crustal Faults:
Extensional Areas; Equal weight for all:
Boore et al., (1997), Sadigh et al. (1997), Abrahamson and Silva (1997), Spudich et al., 1999,
and Campbell and Bozorgnia (2003).
Non-Extensional Areas; Equal weight for all:
Boore et al., (1997), Sadigh et al. (1997), Abrahamson and Silva (1997), and Campbell and
Bozorgnia (2003).
Extensional and non-extensional areas are defined in Figure 5 of the USGS report:
Documentation for the 2002 Update of the National Seismic Hazard Maps, Open-File Report
02-420, U.S. DEPARTMENT OF THE INTERIOR U.S. GEOLOGICAL SURVEY, 2002
Cascadia Subduction Zone:
Youngs et al. (1997) Youngs, R.R., S.J. Chiou, W.J. Silva, and J.R. Humphrey (1997). Strong
ground motion attenuation relationships for subduction zone earthquakes, Seism. Res. Letts., v.
68, no. 1, pp. 58-73.
Sadigh, K., C.Y. Chang, J. Egan, F. Makdisi, and R. Youngs (1997). Attenuation relationships for
shallow crustal earthquakes based on California strong motion data, Seism. Res. Letts., v. 68,
pp. 180-18
Magnitude 9.0:
Use equal weighting for both methods for distances where the Sadigh et al. (1997) PGA values
for M8.5 exceed those of Youngs et al. (1997) for M9.0. For larger distances (R > 60 km), where
the Youngs et al. (1997) PGA values are the higher of the two, use only the Youngs et al. (1997)
relations.
Magnitude 8.3:
Use equal weighting for both methods for distances up to 70km. For distances larger than 70km,
apply full weight to Youngs et al. (1997).
The source distances for the subduction zone events reported from the USGS deaggregation web
site are the closest distances to the fault or slab (R rup). Review the following document for more
information on the proper applications and usage of these attenuation relationships.
Documentation for the 2002 Update of the National Seismic Hazard Maps, Open-File Report
02-420, U.S. DEPARTMENT OF THE INTERIOR U.S. GEOLOGICAL SURVEY, 2002
It is important to note that the ground motion values (PGA, S 0.2, S1.0) obtained for the primary M-R
pairs obtained in this fashion will not likely be the same as the mean values developed for the
Uniform Seismic Hazard (USH), which are used as the basis for structural analysis. Also, it is likely
that the average value of a specific ground motion parameter obtained for the principal M-R pairs will
also vary from the mean value provided by the USGS USH. The difference will reflect the number MR pairs considered and the relative contributions of the sources to the overall hazard.

Volume 1
25

ODOT Geotechnical Design Manual


April 2010

This deaggregation process will likely yield more than one M-R pair, and therefore more than one
magnitude and peak ground acceleration, for liquefaction analysis in some areas of the state where
the hazard is dominated by two or more seismic sources. In most of western Oregon, this will include
both shallow crustal sources and the Cascadia Subduction Zone. In this case, each M-R (i.e., MPGA) pair should be evaluated individually in a liquefaction analysis. If liquefaction is estimated for
any given M-PGA pair, the evaluation of that pair is continued through the slope stability and lateral
deformation evaluation processes. In some areas in the state where the seismic hazard is dominated
by a single source, such as the Cascadia Subduction Zone along the coast, a single pair of M-R
values (largest magnitude (M) and closest distance (R)) may be appropriate for defining and
assessing the worst case liquefaction condition.
The steps involved in a simplified deaggregation application and liquefaction analysis are described
in Dickenson, 2005. Four example problems are provided in Dickenson, 2005 for different areas of
the state, demonstrating the deaggregation procedure. A recommended procedure for estimating
lateral embankment deformations is also included in this paper along with two example problems. A
flow chart of this process, extracted from this paper, is attached as Appendix 6-A.

Site
Characterization
Design

for

Seismic

The geotechnical site investigation should identify and characterize the subsurface conditions and all
geologic hazards that may affect the seismic analysis and design of the proposed structures or
features. The goal of the site characterization for seismic design is to develop the subsurface profile
and soil property information needed for seismic analyses. The geotechnical designer should review
and discuss the project objectives with the project engineering geologist and the structural designer,
as seismic design is a cooperative effort between the geotechnical and structural engineering
disciplines. The geotechnical designer should do the following as a minimum:
Identify potential geologic hazards, areas of concern (e.g., deep soft soils or
liquefiable soils), and potential variability of local geology.

Identify engineering analyses to be performed (e.g., ground response analysis,


liquefaction susceptibility, lateral spreading/slope stability assessments, seismic-induced
settlement/ downdrag, dynamic earth pressures).

Identify engineering properties required for these analyses.

Determine methods to obtain the required design parameters and assess the
validity of such methods for the soil and rock material types.

Develop an integrated investigation of in-situ testing, soil sampling, and laboratory testing. This
includes determining the number of tests/samples needed and appropriate locations to obtain them.

6.4.1

Subsurface Investigation for Seismic Design

Refer to Section 6.0 of AASHTO, 2009, for guidance regarding subsurface investigation and site
characterization for seismic foundation design. With the possible exception of geophysical
explorations associated with obtaining seismic shear wave velocities in soil and rock units, the
subsurface data required for seismic design is typically obtained concurrently with the data required
for static design of the project (i.e., additional exploration for seismic design over and above what is
required for foundation design is typically not necessary). However, the exploration program may
need to be adjusted to obtain the necessary parameters for seismic design. For example, the use of
the seismic cone penetration test, SCPT, is recommended in order to supplement tip resistance and
Volume 1
26

ODOT Geotechnical Design Manual


April 2010

friction data with shear wave velocity. Also, for Site Class determination, subsurface investigations
must extend to a depth of at least 100 feet unless bedrock is encountered before reaching that depth.
The selection of field drilling equipment and sampling methods will reflect the goals of the
investigation. If liquefaction potential is a significant issue, mud rotary drilling with SPT sampling is the
preferred method of investigation. Hollow-stem auger (HSA) drilling may be utilized for SPT sampling
and testing if precautionary measures are taken. Soil heaving and disturbance in HSA borings can
lead to unreliable SPT N values. Therefore care must be taken if using HSA methods to maintain an
adequate water head in the boring at all times and to use drilling techniques that minimize soil
disturbance. Non-standard samplers shall not be used to collect data used in liquefaction analysis
and mitigation design.
In addition to standard subsurface investigation methods, the following equipment calibration, soil
testing, and/or sampling should be considered depending upon site conditions.
SPT Hammer Energy: This value (usually termed hammer efficiency) should be
noted on the boring logs or in the Geotechnical Report. The hammer efficiency should be
obtained from the hammer manufacturer, preferably through field testing of the hammer
system used to conduct the test. This is needed to determine the hammer energy
correction factor, Cer, for liquefaction analysis.

Soil Samples for Gradation Testing: Used for determining the amount
(percentage) of fines in the soil for liquefaction analysis. Also useful for scour estimates.

Undisturbed Samples: Laboratory testing for Su, e50, E, G, and other parameters
for both foundation modeling and seismic design.

Shear Wave Velocity Measurements: For use in determining soil Site Class.
Also used to develop a shear wave velocity profile of the soil column and to obtain low
strain shear modulus values to use in analyses such as dynamic soil response.

Seismic Piezocone Penetrometer: For use in determining soil Site Class. Also
used to develop a shear wave velocity profile and obtain low strain shear modulus
values to use in a ground response analysis.

Piezocone Penetrometer Test: Used for liquefaction analysis and is even


preferred in some locations due to potential difficulties in obtaining good quality SPT
results. Pore pressure measurements and other parameters can be obtained for use in
foundation design and modeling. Also useful in establishing the pre-construction
subsurface soil conditions prior to conducting ground improvement techniques and the
post-construction condition after ground improvement.

Depth to Bedrock: If a ground response analysis is to be performed, the depth


to bedrock must be known. Bedrock material for this purpose is defined as a material
unit with a shear wave velocity of at least 2500 ft/sec.

Pressuremeter Testing: For development of p-y curves if soils cannot be


adequately characterized using standard COM624P or LPile parameters. Testing is
typically performed in soft clays, organic soils, very soft or decomposed rock and for
unusual soil or rock materials. The shear modulus, G, for shallow foundation modeling
and design can also be obtained.

Table 6 -13 provides a summary of site characterization needs and testing considerations for
geotechnical/seismic design.

Volume 1
27

ODOT Geotechnical Design Manual


April 2010

Table 6-13. Summary of site characterization needs and testing considerations for
seismic design (adapted from Sabatini, et al., 2002)

Geotechni
cal Issues

Engineering
Evaluations

Required
Information For
Analyses

Field
Testing

Site Response

source
characterization and
attenuation

SPT

cyclic triaxial tests

CPT

Atterberg Limits

site response
spectra

shear wave
velocity

seismic one

specific gravity

moisture content

time history

bulk shear
modulus for low
strains

geophysical
testing (shear
wave velocity)

unit weight

resonant column

cyclic direct simple


shear test

torsional simple
shear test

subsurface profile
(soil, groundwater,
depth to rock)

piezometer

relationship of
shear modulus
with increasing
shear strain

equivalent viscous
damping ratio with
increasing shear
strain

Poissons ratio

unit weight

relative density

seismicity

Laboratory
Testing

(PGA, design
earthquakes)
Geologic
Hazards
Evaluation (e.g.
liquefaction,
lateral spreading,
slope stability)

liquefaction
susceptibility

liquefaction induced
settlement

settlement of dry
sands

lateral spreading

slope stability and


deformations

subsurface profile
(soil, groundwater,
rock)

SPT

soil shear tests

CPT

seismic cone

triaxial tests
(including cyclic)

Becker
penetration test

grain size
distribution

Atterberg Limits

specific gravity

organic content

moisture content

unit weight

shear strength
(peak and
residual)

unit weights

vane shear test

grain size
distribution

piezometers

geophysical
testing (shear
wave velocity)

plasticity
characteristics

relative density

penetration
resistance

shear wave
velocity

seismicity (PGA,
design
earthquakes)

site topography

Volume 1
28

ODOT Geotechnical Design Manual


April 2010

Table 6-1 Summary of site characterization needs and testing considerations for seismic
design (contd) (adapted from Sabatini, et al., 2002).

Geotechni

cal

Engineering

Required

Evaluations

Information

For Analyses

Issues

Input for
Structural Design

shallow foundation
springs

p-y data for deep


foundations

down-drag on deep
foundations

residual strength

lateral earth
pressures

lateral spreading/
slope movement
loading
post earthquake
settlement

Field
Testing

subsurface profile
(soil, groundwater,
rock)

CPT

triaxial tests

SPT

soil shear tests

shear strength
(peak and
residual)

seismic cone

piezometers

unconfined
compression

geophysical
testing (shear
wave velocity)

grain size
distribution

Atterberg Limits

specific gravity

moisture content

unit weight

resonant column

cyclic direct simple


shear test

torsional simple
shear test

seismic horizontal
earth pressure
coefficients
shear modulus for
low strains or
shear wave
velocity
relationship of
shear modulus
with increasing
shear strain

unit weight

Poissons ratio

seismicity (PGA,
design
earthquake)

site topography

vane shear test

Laboratory
Testing

For routine designs, in-situ or laboratory testing for parameters such as the dynamic shear modulus
at small strains, equivalent viscous damping, shear modulus and damping ratio versus shear strain,
and residual shear strength are generally not directly obtained. Instead, index properties and
correlations based on in-situ field measurements (such as the SPT and CPT) are generally used in
lieu of in-situ or laboratory measurements for routine design to estimate these values.
If correlations are used to obtain seismic soil design properties, the following correlations are
recommended. Other acceptable correlations can be found in Dickenson et al. (2002), Kramer
(1996), and other technical references. Region and site-specific correlations developed by
practitioners are acceptable with adequate supporting documentation and approval by ODOT.
ODOT Table 6 -14, which presents correlations for estimating initial shear
modulus (Gmax) based on relative density, penetration resistance or void ratio.

ODOT Figure 6 -2, which presents shear modulus reduction curves and
equivalent viscous damping ratio for cohesionless soils (sands) as a function of shear
strain and depth.

ODOT Figure 6 -3 and Figure 6 -4, which present shear modulus reduction
curves and equivalent viscous damping ratio, respectively, as a function of cyclic shear
strain and plasticity index for fine grained (cohesive) soils.

Volume 1
29

ODOT Geotechnical Design Manual


April 2010

ODOT Figure 6 -5, Figure 6 -6 and Figure 6 -7 which presents a chart for
estimating undrained residual shear strength for liquefied soils as a function of SPT blow
counts (N60) and vertical effective stress.

Table 6-14. Correlations for estimating initial shear modulus (Kavazajjian, et al., 1997).
Chapter 15 Reference

Chapter 25 Correlation

Chapter 35 U Chapter 45 Limitations

nits (1)

Seed et al.

(1984)

Gmax = 220 (K2)max (m)

kPa

Imai and
Tonouchi

(1982)

Mayne and Rix

(1993)

Notes:

(K2)max is about 30 for


very loose sands and 75 for
very dense sands; about 80
to 180 for dense well graded
gravels; Limited to
cohesionless soils

(K2)max = 20(N1)60

Gmax = 15,560 N600.68

kPa

Gmax = 99.5(Pa)0.305(qc)0.695/(e0)1.13

kPa(

(1) 1 kPa = 20.885 psf

(2) Pa and qc in kPa

Volume 1
30

Limited to
cohesionless soils

2)

Limited to cohesive
soils; Pa = atmospheric
pressure

ODOT Geotechnical Design Manual


April 2010

Figure 6-2. Shear modulus reduction and damping ratio curves for sand (EPRI, 1993).

Volume 1
31

ODOT Geotechnical Design Manual


April 2010

Figure 6-3. Variation of G/Gmax vs. cyclic shear strain for fine grained soils (redrafted
from Vucetic and Dobry, 1991).

Figure 6-4. Equivalent viscous damping ratio vs. cyclic shear strain for fine grained soils
(redrafted from Vucetic and Dobry, 1991).

Volume 1
32

ODOT Geotechnical Design Manual


April 2010

Figure 6-5. Residual undrained shear strength for liquefied soils as a function of SPT
blow counts (Seed and Harder, 1990 and Idriss, 2003).

Figure 6-6. Estimation of residual strength ratio from SPT resistance (Olson and Stark,
2002).

Volume 1
33

ODOT Geotechnical Design Manual


April 2010

Figure 6-7. Variation of residual strength ratio with SPT resistance and initial vertical
effective stress using Kramer-Wang model (Kramer, 2008).

Geotechnical
Procedures

Seismic

Design

The geotechnical designer shall evaluate the site and subsurface conditions to the extent necessary
to provide the following assessments and recommendations:

an assessment of the seismic hazard,

determination of design ground motion values,

site characterization,

seismic analysis of the foundation materials and

an assessment of the effects of the foundation response on the proposed


structure.

Specific aspects of seismic foundation design generally consist of the following procedures:
Determine the Peak Bedrock Acceleration (PGA), 0.2 and 1.0 second spectral
accelerations for the bridge site from the 2002 USGS National Seismic Hazard Maps for
the 500 and 1000-year return periods,

Determine the Site Class and Site Coefficients based on the properties of the soil

profile,

Volume 1
34

ODOT Geotechnical Design Manual


April 2010

Develop the Design Response Spectrum for the site or conduct ground response
analysis if necessary,

Determine liquefaction potential of foundation soils,

If liquefaction is predicted:
1. Estimate embankment deformations (lateral spread), bridge damage potential and
approach fill performance for both the 500 and 1000-year events.
2. Determine seismic fill settlement (potential downdrag and bridge damage if
applicable).
3. Provide soil properties for both the liquefied and non-liquefied soil conditions for use
in the lateral load analysis of deep foundations.
4. Determine reduced foundation resistances and their effects on proposed bridge
foundation elements.

Evaluate slope stability and settlement for non-liquefied soil conditions.

Evaluate impacts of seismic geologic hazards including liquefaction, lateral


spreading and slope instability on infrastructure, including estimated loads and
deformations acting on the structure.

Develop foundation spring values for dynamic loading (liquefied and nonliquefied soil conditions). Also recommendations regarding lateral springs for use in
modeling abutment backfill soil resistance.

Determine earthquake induced earth pressures (active and passive) and provide
stiffness values for equivalent soil springs (if required) for retaining structures and below
grade walls.

Evaluate options to mitigate seismic geologic hazards, such as ground


improvement, if appropriate.

Note that separate analysis and recommendations will be required for the 500 and 1000 year seismic
design ground motions. A general design procedure is described in the flow chart shown in Figure 6
-8 along with the information that should be supplied in the final geotechnical report.

Volume 1
35

ODOT Geotechnical Design Manual


April 2010

STEP 1; Ground Motion Data


Identify the seismic sources in the region affecting the site for the given return period (500 and 1000 yrs):
Determine Peak Ground Accelerations (PGA), Ss, and S1, from the 2002 USGS Seismic Hazard Maps for bedrock (Site
Class B) conditions
Determine Site Class and Site Coefficients, then develop the Design Response Spectrum representing the Uniform
Seismic Hazard

STEP 2; Site Response Analysis


Decide whether a site response analysis is warranted.
If so, perform deaggregation of seismic hazard to determine principal M & R pairs.
Select appropriate acceleration time histories and establish scaling factors or perform spectral matching.
Generate the following:
PGA and 5% damped smoothed response spectra at the depth(s) of interest (e.g., ground surface, depth of
pile/pier fixity).
Profiles of PGA, cyclic shear strain, and cyclic shear stress with depth.

STEP 3; Evaluate Liquefaction Potential & Effects (PGA0.10g)


Estimate the cyclic resistance of the soils as a function of depth from in situ and/or lab data.
Specify the cyclic loading at each depth from Step 2.
Using the ratio of the cyclic resistance to the cyclic loading determine the potential for significant excess pore
pressure generation and cyclic degradation of soil stiffness and strength.

No Liquefaction Potential

Liquefaction Potential

STEP 3a; For foundation soils susceptible to


liquefaction:
Estimate post-liquefaction soil strengths
Evaluate embankment stability and est. deformations
Develop mitigation designs if required
Assess the effects of liquefaction on foundation resistances
and provide reduced foundation resistances under liquefied
soil conditions. (CHECK DOWNDRAG)

STEP 3a; Evaluate Non-liquefied Soil


Response
Dynamic settlement of foundation soils and downdrag
potential
Evaluate approach fill slope stability
Estimate lateral approach fill displacements

STEP 4; Provide seismic foundation modeling parameters


as appropriate (see Section 1.1.4 of BDDM):

Spread Footings
Shear modulus; G is dependent on the
shear strain; generally a G corresponding
to a shear strain in the range of 0.20% to
0.02% is appropriate. For large magnitude
events (M>7.5) and very high PGA
(>0.6g), a G corresponding to a shear
strain of 1% is recommended. A ground
response analysis may also be conducted
to determine the appropriate shear strain
value to use.
Poisons ratio,
Kp, Su, ,

Piles
p-y curve data for nonliquefied and liquefied soils
p-y multipliers
Designation as end bearing
or friction piles for modeling
axial stiffness

Shafts
p-y curve data for non-liquefied
and liquefied soils
p-y multipliers

Figure 6-8. General Geotechnical Seismic Design Procedures

Volume 1
36

ODOT Geotechnical Design Manual


April 2010

6.5.1

Design Ground Motion Data

6.5.1.1

D e ve l o p m e n t o f D e s i g n G r o u n d M o t i o n D at a

In general, there are two options for the development of design ground motion parameters (response
spectral ordinates) for seismic design. Both procedures are based on the USGS 2002 PSHA maps.
These are described as follows:
1. AASHTO General Procedure: Use specification/code based hazard (2002 USGS Maps)
with specification/code based site coefficients.
2. Ground Response Analysis: Use specification/code based hazard (2002 USGS Maps) with
site specific ground response analysis.
Both methods take local site effects into account. For most routine structures at sites with competent
soils (i.e., no liquefiable, sensitive, or weak soils), the first method (General Procedure), described in
Article 3.4 of the AASHTO Guide Specification for LRFD Seismic Bridge Design, is sufficient to
account for site effects. However, the importance of the structure, the ground motion levels and the
soil and geological conditions of a site may dictate the need for a Ground Response Analysis
(second method). The geotechnical engineer is responsible for developing and providing the design
response spectra for the project.

6.5.1.2

AAS H T O G e n e r a l P r o c e d u r e

The standard method of developing the acceleration response spectrum is described in AASHTO,
2009. First, the peak ground acceleration (PGA), the short-period spectral acceleration (S s) and the
long-period spectral acceleration (S1) are obtained from the 2002 USGS Seismic Hazard Maps for
the location of the bridge. PGA, Ss, and S1 are obtained for both the 500-year and 1000-year return
periods. Then the soil profile is classified as one of six different site classes (A through F). This Site
Class designation is then used to determine the Site Coefficients, Fpga, Fa and Fv, except for sites
classified as Site Class F, which required a site-specific ground response analysis Section 6.5.3.
These site coefficients are then multiplied by the peak ground acceleration (Fpga x PGA), the shortperiod spectral acceleration (Fa x Ss) and the long period spectral acceleration (Fv x S1) respectively
and used to develop the site response spectrum. A program to develop the response spectra using
the general procedure has been developed by the Bridge Section and can be accesses through the
ODOT Bridge Section web page.
Once the response spectrum is developed the structural engineer can determine the Response
Spectral Acceleration (per AASHTO 2009 Guide Specifications) for use in the seismic design of the
structure.

6.5.1.3

R e s p o n s e S p e ct r a an d An a l ys i s f o r L i q u ef i ed
Soil Sites

Site coefficients have not been developed for liquefied soil conditions. For this case site-specific
analysis is required to estimate ground motion characteristics. The AASHTO Guide Specifications for
LRFD Seismic Bridge Design states that at sites where soils are predicted to liquefy the bridge shall
be analyzed and designed under two configurations, the nonliquefied condition and liquefied soil
condition described as follows:

Volume 1
37

ODOT Geotechnical Design Manual


April 2010

Nonliquefied Configuration: The structure is analyzed and designed, assuming


no liquefaction occurs by using ground response spectrum and soil design parameters
based on nonliquefied soil conditions.

Liquefied Configuration: The structure is reanalyzed and designed under


liquefied soil conditions assuming the appropriate residual resistance for lateral and axial
deep foundation response analyses consistent with liquefied soil conditions (i.e.,
modified P-Y curves, modulus of subgrade reaction, T-Z curves, axial soil frictional
resistance). The design spectrum should be the same as that used in nonliquefied
configuration.

A site-specific response spectrum may be developed for the Liquefied Configuration based on a
ground response analysis that utilizes non-linear, effective stress methods, which properly account
for pore pressure buildup and stiffness degradation of the liquefiable soil layers see Section 6.5.1.4.
The decision to complete a ground response analysis where liquefaction is anticipated should be
made by the geotechnical designer based on the site geology and characteristics of the bridge being
designed. The design response spectrum resulting from the ground response analyses shall not be
less than two-thirds of the spectrum developed using the general procedure for the non-liquefied soil
condition.

6.5.1.4

G r o u n d R e s p o n s e An a l ys i s

For most projects, the General Procedure as described in Article 3.4.1 of the AASHTO Guide
Specifications for LRFD Seismic Bridge Design is appropriate and sufficient for determining the
seismic hazard and site response spectrum. However, it may be appropriate to perform a sitespecific evaluation for cases involving special aspects of seismic hazard (e.g., near fault conditions,
high ground motion values, coastal sites located in relatively close proximity to the CSZ source),
specific soil profiles, and essential bridges. The results of the site-specific response analysis may be
used as justification for a reduction in the spectral response ordinates determined using the standard
AASHTO design spectrum (General Procedure) representing the Uniform Seismic Hazard.
Site specific ground response analyses (GRA) are required for Site Class F soil profiles, and may
be warranted for other site conditions or project requirements. Site Class F soils are defined as
follows:

Peat or highly organic clays, greater than 10 ft in thickness,

Very high plasticity clays (H > 25 ft with PI > 75),

Very thick soft/medium stiff clays (H >120 ft),

Other conditions under which a ground response analysis should be considered are listed below:

Very important or critical structures or facilities,

Liquefiable Soil Conditions. For liquefiable soil sites, it may be desirable to


develop response spectra that take into account increases in pore water pressure and
soil softening. This analysis results in a response spectra that is generally lower than the
nonliquefied response spectra except for spectral accelerations in the higher period
range (above 1.0 second). A nonlinear effective stress analysis may also be necessary
to refine the standard liquefaction analysis based on Seeds Simplified (SPT) Method (or
others) with information from a GRA. This is especially true if liquefaction mitigation
designs are proposed. The cost of liquefaction mitigation is sometimes very large and a
more detailed analysis to verify the potential, and extent, of liquefaction is usually
warranted.

Volume 1
38

ODOT Geotechnical Design Manual


April 2010

very deep soil deposits or thin (<40 50 feet) soil layers over bedrock.

to obtain better information for evaluating lateral deformations, near surface soil
shear strain levels or deep foundation performance.

to obtain ground surface PGA values for abutment wall or other design.

Procedures for conducting a site specific ground response analysis are described in Article 3.4.3. of
the AASHTO guide specifications and in Chapter 4 of Kavazanjian, et al. (1997).
A ground response analysis evaluates the response of a layered soil deposit subjected to earthquake
motions. One-dimensional, equivalent-linear models are commonly utilized in practice. This model
uses an iterative total stress approach to estimate the nonlinear elastic behavior of soils. Modified
versions of the numerical model SHAKE (e.g., ProSHAKE, SHAKE91, SHAKE2000) are routinely
used to simulate the propagation of seismic waves through the soil column and generate output
consisting of ground motion time histories at selected locations in the soil profile, plots of ground
motion parameters with depth (e.g., PGA, cyclic shear stress, cyclic shear strain), and acceleration
response spectra at depths of interest. The program calculates the induced cyclic shear stresses in
individual soil layers which may be used in liquefaction analysis.
The equivalent linear model provides reasonable results for small to moderate cyclic shear strains
(less than about 1 to 2 percent) and modest accelerations (less than about 0.3 to 0.4g) (Kramer and
Paulsen, 2004). Equivalent linear analysis cannot be used where large strain incompatibilities are
present, to estimate permanent displacements, or to model development of pore water pressures in
a coupled manner. Computer programs capable of modeling non-linear, effective stress soil behavior
are recommended for sites where high ground motion levels are indicated and it is anticipated that
moderate to large shear strains will be mobilized. These are typically sites with soft to medium stiff
fine-grained soils or saturated deposits of loose to medium dense cohesionless soils.
Basically, the input parameters required for site specific seismic response analysis include soil
layering (thickness), standard geotechnical index properties for the soils, dynamic soil properties for
each soil layer, the depth to bedrock or firm soil interface, and a set of ground motion time histories
representative of the primary seismic hazards in the region.
Soil parameters required by the equivalent linear models include the shear wave velocity or initial
(small strain) shear modulus and unit weight for each soil layer, and curves relating the shear
modulus and damping ratio as a function of shear strain (see Section 6.4.1 and Figure 6 -2, Figure
6 -3 and Figure 6 -4 for examples).

6.5.1.5

S e l e c t i o n o f Tim e H i st o r i e s f o r G r o u n d
R e s p o n s e An a l ys i s

AASHTO (2009) allows two options for the selection of time histories to use in ground response
analysis. The two options are:
a) Use a suite of 3 response-spectrum-compatible time histories with the design response
spectrum developed enveloping the maximum response, or
b) Use of at least 7 time histories and develop the design spectrum as the mean of the
computed response spectra.
For both options, the time histories shall be developed from the representative recorded earthquake
motions, or in special instances synthetic ground motions may be used with approval of ODOT. The
time histories for these applications shall have characteristics that are representative of the seismic
environment of the site and the local site conditions, including the response spectrum for the site.

Volume 1
39

ODOT Geotechnical Design Manual


April 2010

Analytical techniques used for spectral matching shall be demonstrated to be capable of achieving
seismologically realistic time series. The time histories should be scaled to the approximate level of
the design response spectrum in the period range of significance (i.e., 0.5 < T < 2.0).
The procedures for selecting and scaling time histories for use in ground motion response analysis
can be summarized as follows:
1. Identify the target response spectra to be used to develop the time histories. The target
spectra are obtained from the 2002 USGS Seismic Hazard Maps for top-of-rock locations
(base of soil column). Two spectra are required, one for the 500-yr return event and one for
the 1000-yr event.
2. Identify the seismic sources that contribute to the seismic hazard for the site, considering the
desired probability of exceedance (i.e., 500 and 1000-yr return periods). Use the
deaggregation information for the 2002 USGS Seismic Hazard maps to obtain information on
the primary sources that affect the site. All seismic sources (M-R pairs) that contribute more
than 5% to the hazard in the period range of interest should be considered.
3. Select time histories to be considered for the analysis, considering tectonic environment and
style of faulting (subduction zone, Benioff zone, or shallow crustal faults), seismic source-tosite-distance, earthquake magnitude, duration of strong shaking, peak acceleration, site
subsurface characteristics, predominant period, etc. In areas where the hazard has a
significant contribution from both the Cascadia Subduction Zone (CSZ) and from crustal
sources (e.g., Portland and much of the Western part of the state) both earthquake sources
need to be included in the analysis and development of a site specific response spectra. In
cases such as this, it is recommended that the ground response analysis be conducted using
a collection of time histories that include at least 3 motions representative of subduction zone
events and 3 motions appropriate for shallow crustal earthquakes with the design response
spectrum developed considering the mean spectrum of each of these primary sources.

Volume 1
40

ODOT Geotechnical Design Manual


April 2010

At sites where the uniform hazard is dominated by a single source, three (3) time histories,
representing the seismic source characteristics, may be used and the design response spectrum
determined by enveloping the caps of the resulting response spectra.
4. Scale the time histories to match the target spectrum as closely as possible in the period
range of interest prior to spectral matching. Match the response spectra from the recorded
earthquake time histories to the target spectra using methods that utilize either time series
adjustments in the time domain or adjustments made in the frequency domain. See
AASHTO Guide Specifications for LRFD Seismic Bridge Design and Kramer (1996) for
additional guidance on these techniques.
5. Once the time history(ies) have been spectrally matched, they can be used directly as input
into the ground response analysis programs to develop response spectra and other seismic
design parameters. Five percent (5%) damping is typically used in all site response analysis.
The results of the dynamic soil response modeling should be presented as the mean, average
and 85th percentile curves from all of the output response spectra. A smoothed response spectra
may be obtained by enveloping the peaks of the 85% percentile response curve. The resulting
design spectrum may be lower than the 85th percentile curve outside of the period range of interest.
Engineering judgment will be required to account for possible limitations of the response modeling.
For example, equivalent linear analysis methods may overemphasize spectral response where the
predominant period of the soil profile closely matches the predominant period of the bedrock motion.
Final modification of the design spectrum must provide representative constant velocity and constant
displacement portions of the response. Site-specific response spectra may be used for design
however the lower limit shall be no less than 2/3rd of the AASHTO response spectrum using the
General Procedure. An example response spectrum is attached in Appendix 6-B.
At some bridge sites, the subsurface conditions (soil profile) may change dramatically along the
length of the bridge and more than one response spectrum may be required to represent segments
of the bridge with different soil profiles. If the site conditions dictate the need for more than one
response spectrum for the bridge, the design response spectrum should be developed by combining
the individual spectra into a composite spectrum that envelopes the spectral acceleration values of
the individual spectra.
Nonlinear effective stress analysis methods such as D-MOD, DESRA and others may be used to
develop response spectra especially at sites where liquefaction of foundation soils is likely. All nonlinear, effective stress modeling and analysis will require an independent peer reviewer with expertise
in this type of analysis. In some cases, the response spectra resulting from a nonlinear effective
stress analysis may result in spectral acceleration values exceeding the 2/3 AASHTO general
procedure criteria in the range of higher periods. If this is the case, the higher response spectra
values of the two methods shall be used. For non-linear analysis methods the lower limit shall be no
less than 2/3rd of the AASHTO response spectrum developed using the General Procedure.

6.5.1.6

Ground Motion Parameters for Other Structures

For buildings, restrooms, shelters, and other non-transportation structures, specification based
seismic design parameters required by the 2003 IBC should be used. The seismic design
requirements of the 2003 IBC are based on a risk level of 2 percent PE in 50 years. The 2 percent
PE in 50 years risk level corresponds to the maximum considered earthquake. The 2003 IBC
identifies procedures to develop a maximum considered earthquake acceleration response
spectrum, and defines the design response spectrum as two-thirds of the value of the maximum
considered earthquake acceleration response spectrum.

Volume 1
41

ODOT Geotechnical Design Manual


April 2010

Site response shall be in accordance with the 2003 IBC. As is true for transportation structures, for
critical or unique structures or for sites characterized as soil profile Type F (thick sequence of soft
soils or liquefiable soils), site response analysis may be required.

6.5.1.7

B e d r o c k ve r s u s G ro u n d S u r f a ce Ac c e l e r a t i o n

Soil amplification factors that account for the presence of soil over bedrock with regard to the
estimation of peak ground acceleration (PGA) are directly incorporated into the development of the
general procedure for developing response spectra for structural design of bridges and similar
structures in the AASHTO LRFD Bridge Design and Guide Specifications and for the structural
design of buildings and non-transportation related structures in the 2003 IBC. Additional amplification
factors should not be applied to peak bedrock accelerations when code based response spectra are
used. However, amplification factors should be applied to the peak bedrock acceleration to determine
the peak ground acceleration (PGA) for liquefaction assessment, such as for use with the Simplified
Method Section 6.5.5.2 , and for the estimation of seismic earth pressures and inertial forces for
retaining wall and slope design. For liquefaction assessment and retaining wall and slope design, the
Site Factors (Fpga) presented in AASHTO 3.10.3.2 may be applied to the bedrock PGA used to
determine the ground surface acceleration, unless a site specific evaluation of ground response is
conducted.

6.5.2

Liquefaction Analysis

Liquefaction has been one of the most significant causes of damage to bridge structures during past
earthquakes. Liquefaction can damage bridges and structures in many ways including:

Bearing failure of shallow foundations founded above liquefied soil;

Liquefaction induced ground settlement;

Lateral spreading of liquefied ground;

Large displacements associated with low frequency ground motion;

Increased earth pressures on subsurface structures;

Floating of buoyant, buried structures; and

Retaining wall failure.

Liquefaction refers to the significant loss of strength and stiffness resulting from the generation of
excess pore water pressure in saturated, cohesionless soils. Liquefaction can occur in sand and nonplastic to low plasticity silt-rich soils, and in confined gravel layers; however, it is most common in
sands and silty sands. For a detailed discussion of the effects of liquefaction, including the types of
liquefaction phenomena, liquefaction-induced bridge damage, evaluation of liquefaction susceptibility,
post liquefaction soil behavior, deformation analysis and liquefaction mitigation techniques refer to
Dickenson, et al. (2002).
Liquefaction hazard assessment includes identifying soils susceptible to liquefaction on the basis of
composition and cyclic resistance, evaluating whether the design earthquake loading will initiate
liquefaction, and estimating the potential effects of liquefaction on the planned facility. Potential
effects of liquefaction on soils and foundations include the following:

Loss in strength in the liquefied layer(s)

Liquefaction-induced ground settlement

Flow failures, lateral spreading, and slope instability.

Volume 1
42

ODOT Geotechnical Design Manual


April 2010

Due to the high cost of liquefaction mitigation measures, it is important to identify liquefiable soils and
the potential need for mitigation measures early on in the design process (during the DAP (TS&L)
phase) so that appropriate and adequate funding decisions are made. The following sections provide
ODOTs policies regarding liquefaction and a general overview of liquefaction hazard assessment
and its mitigation. .

6.5.2.1

Liquefaction Design Policies

All new bridges, bridge widening projects and retaining walls in areas with seismic acceleration
coefficients, or PGA, greater than or equal to 0.10g should be evaluated for liquefaction potential. The
maximum considered liquefaction depth shall be limited to 75 feet. The potential for liquefaction and
limited strength and stiffness reductions due to pore pressure increase caused by ground shaking
may be considered below this depth on the basis of cyclic laboratory test data and/or the use of nonlinear, effective stress analysis techniques. All non-linear, effective stress modeling and analysis will
require an independent peer reviewer with expertise in this type of analysis.
Bridges scheduled for seismic retrofit should also be evaluated for liquefaction potential if they are in
a seismic zone with an acceleration coefficient (or PGA) 0.10g.
In general, liquefaction is conservatively predicted to occur when the factor of safety against
liquefaction (FSL) is less than 1.1. A factor of safety against liquefaction of 1.1 or less also indicates
the potential for liquefaction-induced ground movement (lateral spread and settlement). Soil layers
with FSL between 1.1 and 1.4 will have reduced soil shear strengths due to excess pore pressure
generation. For soil layers with FSL greater than 1.4, excess pore pressure generation is considered
negligible and the soil does not experience appreciable reduction in shear strength.
If liquefaction is predicted based on the Simplified Method Section 6.5.2.2, and the effects of
liquefaction require mitigation measures, a more thorough ground response analysis (e.g. SHAKE,
DMOD) is recommended to verify and substantiate the predicted, induced ground motions. This
procedure is especially recommended for sites where liquefaction potential is marginal (0.9 < FS L <
1.10). It is also important to determine whether the liquefied soil layer is stratigraphically (laterally)
continuous and oriented in a manner that will result in lateral spread or other adverse impact to the
bridge.
Groundwater: The groundwater level to use in the liquefaction analysis should be determined as
follows:

Static Groundwater Condition: Use the estimated, average annual groundwater level.
Perched water tables should only be used if water is estimated to be present in these
zones more than 50% of the year.

Tidal Areas: Use the mean high tide elevation.

Adjacent Stream, Lake or Standing Water Influence: Use the estimated, annual,
average elevation for the wettest (6 month) seasonal period.

Volume 1
43

ODOT Geotechnical Design Manual


April 2010

Note that groundwater levels measured in borings advanced using water or other drilling fluids may
not be indicative of true static groundwater levels. Water in these borings should be allowed to
stabilize over a period of time to insure measured levels reflect true static groundwater levels.
Groundwater levels are preferably measured and monitored using piezometers, taking
measurements throughout the climate year to establish reliable static groundwater levels taking
seasonal effects into account.

6.5.2.2

M e t h o d s t o E va l u a t e L i q u ef a ct i o n P o t en t i a l

Evaluation of liquefaction potential should be based on soil characterization using in-situ testing
methods such as Standard Penetration Tests (SPT) and Cone Penetration Tests (CPT). Liquefaction
potential may also be evaluated using shear wave velocity (Vs) testing and Becker Penetration Tests
(BPT); however, these methods are not preferred and are used less frequently than SPT or CPT
methods. Vs and BPT testing may be appropriate in soils difficult to test using SPT and CPT methods
such as gravelly soils though, in the absence of fine grained soil layers that may act as poorly drained
boundaries, these soils often have a low susceptibility to liquefaction potential due to high
permeability and rapid drainage. If the CPT method is used, SPT sampling and soil gradation testing
shall still be conducted to obtain direct information on soil type and gradation parameters for use in
liquefaction susceptibility assessment.
Preliminary Screening: A detailed evaluation of liquefaction potential is not required if any of the
following conditions are met:

The bedrock PGA (or Acceleration Coefficient, As) is less than 0.10g,

The ground water table is more than 75 feet below the ground surface,

The soils in the upper 75 feet of the profile have a minimum SPT resistance, corrected for
overburden depth and hammer energy (N160), of 25 blows/ft, or a cone tip resistance qciN of
150 tsf.

All soils in the upper 75 feet are classified as cohesive, and have a PI 18. Note that
cohesive soils with PI 18 may still be very soft or exhibit sensitive behavior and could
therefore undergo significant strength loss under earthquake shaking. This criterion should be
used with care and good engineering judgment. Recent advances in the screening and
evaluation of fine-grained soils for strength loss during cyclic loading can be found in Bray
and Sancio, (2006) and Boulanger and Idriss, (2006, 2007).

Simplified Procedures: Simplified Procedures should always be used to evaluate the liquefaction
potential even if more rigorous methods are used to supplement or refine the analysis. The Simplified
Procedure was originally developed by Seed and Idriss (1971) and has been periodically modified
and improved since. It is routinely used to evaluate liquefaction resistance in geotechnical practice.
The procedures described in Section 3.4 of the report Assessment and Mitigation of Liquefaction
Hazards to Bridge Approach Embankments in Oregon, (Dickenson et al, 2003) should be followed
for assessing the liquefaction potential of soil by the Simplified Procedures.
The paper titled Liquefaction Resistance of Soils: Summary Report from the 1996 NCEER/NSF
Workshops on Evaluation of Liquefaction Resistance of Soils (Youd et al., (2001) also provides a
state of the practice summary of the Simplified Procedures for assessment of liquefaction
susceptibility. This paper resulted from a 1996 workshop of liquefaction experts sponsored by the
National Center for Earthquake Engineering Research and the National Science Foundation with the
Volume 1
44

ODOT Geotechnical Design Manual


April 2010

objective being to gain consensus on updates and augmentation of the Simplified Procedures. Youd
et al. (2001) provide procedures for evaluating liquefaction susceptibility using SPT, CPT, Vs, and
BPT criteria.
The Simplified Procedures are based on the evaluation of both the cyclic resistance ratio (CRR) of a
soil layer (i.e., the cyclic shear stress required to cause liquefaction) and the earthquake induced
cyclic shear stress ratio (CSR). The resistance value (CRR) is estimated based on empirical charts
relating the resistance available to specific index properties (i.e. SPT, CPT, BPT or shear wave
velocity values) and corrected to an equivalent magnitude of 7.5 using a magnitude scaling factor.
Youd et al. (2001) provide the empirical liquefaction resistance charts for both SPT and CPT data to
be used with the simplified procedures.
The basic form of the simplified procedures used to calculate the earthquake induced CSR for the
Simplified Method is shown in the following equation:

Equation 6.1

Where:

T
av

= average or uniform earthquake induced cyclic shear


stress

amax = peak horizontal acceleration at the ground surface accounting


for site amplification effects (ft/sec2)
g = acceleration due to gravity (ft/sec2)
o = initial total vertical stress at depth being evaluated (lb/ft2)
o = initial effective vertical stress at depth being evaluated (lb/ft2)
rd = stress reduction coefficient
The factor of safety against liquefaction is defined by:
FSliq = CRR/CSR
The use of the SPT for the Simplified Procedure has been most widely used and has the advantage
of providing soil samples for fines content and gradation testing. The CPT provides the most detailed
soil stratigraphy, is less expensive, can simultaneously provide shear wave velocity measurements,
and is more reproducible. If the CPT is used, soil samples shall be obtained using the SPT or other
methods so that detailed gradational and plasticity analyses can be conducted. The use of both SPT
and CPT procedures can provide the most detailed liquefaction assessment for a site.
Where SPT data is used, the sampling and testing procedures should include:

Documentation on the hammer efficiency (energy measurements) of the system

used.
Correction factors for borehole diameter, rod length and sampler liners should be
used, where appropriate.

Where gravels or cobbles are present, the use of short interval adjusted SPT N
values may be effective for estimating the N values for the portions of the sample not
affected by gravels or cobbles.

Volume 1
45

ODOT Geotechnical Design Manual


April 2010

Blowcounts obtained using non-standard samplers such as the Dames and


Moore or modified California samplers shall not be used for liquefaction evaluations.

Limitations of the Simplified Procedures: The limitations of the Simplified Procedures should be
recognized. The Simplified Procedures were developed from empirical evaluations of field
observations. Most of the case history data was collected from level to gently sloping terrain
underlain by Holocene-age alluvial or fluvial sediment at depths less than 50 feet. Therefore, the
Simplified Procedures are applicable to only these site conditions. Caution should be used for
evaluating liquefaction potential at depths greater than 50 feet using the Simplified Procedure. In
addition, the Simplified Procedures estimate the trend of earthquake induced cyclic shear stress ratio
with depth based on a coefficient, rd, which becomes highly variable at depths below about 40 feet.
As an alternative to the use of Equation 6.1, one dimensional ground response analyses should be
used to better determine the maximum earthquake induced shear stresses at depths greater than
about 50 feet. Equivalent linear, total stress computer programs (Shake2000, ProShake or other
equivalent program) may be used for this purpose.
Nonlinear Effective Stress Methods: An alternative to the simplified procedures for evaluating
liquefaction susceptibility is to perform a nonlinear, effective stress site response analysis utilizing a
computer code capable of modeling pore water pressure generation and dissipation (D-MOD2000,
DESRA, FLAC). These are more rigorous analyses and they require additional soil parameters,
validation by the practitioner, and additional specialization.
The advantages of this method of analysis include the ability to assess liquefaction at depths greater
than 50 feet, the effects of liquefaction and large shear strains on the ground motion, and the effects
of higher accelerations that can be more reliably evaluated. In addition, seismically induced
deformation can be estimated, and the timing of liquefaction and its effects on ground motion at and
below the ground surface can be assessed.
Several non-linear, effective stress analysis programs can be used to estimate liquefaction
susceptibility at depth. However, few of these programs are being used by geotechnical designers at
this time. In addition, there has been little verification of the ability of these programs to predict
liquefaction at depths greater than 50 feet because there are few well documented sites of deep
liquefaction.
Due to the highly specialized nature of these more sophisticated liquefaction assessment
approaches, an independent peer review by an expert in this type of analysis is required to use
nonlinear effective stress methods for liquefaction evaluation.
Magnitude and PGA for Liquefaction Analysis: The procedures described in Section 6.3.2, and
in Dickenson et al. (2002), should be used to determine the appropriate earthquake magnitude and
peak ground surface acceleration to use in the simplified procedure for liquefaction analysis. If a site
specific ground response analysis is used to determine the peak ground surface acceleration(s) for
use in liquefaction analyses, this value should be representative of the cyclic loading induced by the
M-R pair(s) of interest. It is anticipated that PGA values obtained from site-specific ground response
analysis will differ from the PGA determined by the AASHTO General Procedure for the Uniform
Seismic Hazard. The PGA and magnitude values used in the liquefaction hazard analysis shall be
tabulated for all considered seismic sources.
Magnitude Scaling Factors (MSF): Magnitude scaling factors are required to adjust the cyclic
stress ratios (either CRR or CSR) obtained from the Simplified Method (based on M = 7.5) to other
magnitude earthquakes. The range of Magnitude Scaling Factors recommended in the 1996 NCEER
Workshop on Evaluation of Liquefaction Resistance of Soils (Youd, et. al., 2001) is recommended.
Below magnitude 7.5, a range is provided and engineering judgment is required for selection of the
Volume 1
46

ODOT Geotechnical Design Manual


April 2010

MSF. Factors more in line with the lower bound range of the curve are recommended. Above
magnitude 7.5 the factors recommended by Idriss are recommended. This relationship is presented
in the graph (Figure 6 -9) and the equation of the curve is: MSF = 102.24 / M2.5

Figure 6-9. Magnitude Scaling Factors Derived by Various Investigators (redrafted from
1996 NCEER Workshop Summary Report)

6.5.2.3

Liquefaction Induced Settlement

Both dry and saturated deposits of loose granular soils tend to densify and settle during earthquake
shaking. Settlement of unsaturated (dry) granular deposits is discussed in Section 6.5.4. If the
Simplified Procedure is used to evaluate liquefaction potential, liquefaction induced ground
settlement of saturated granular deposits should be estimated using the procedures by Tokimatsu
and Seed (1987) or Ishihara and Yoshimine (1992). The Tokimatsu and Seed (1987) procedure
estimates the volumetric strain as a function of earthquake induced CSR and corrected SPT
blowcounts. The Ishihara and Yoshimine (1992) procedure estimates the volumetric strain as a
function of factor of safety against liquefaction, relative density, and corrected SPT blowcounts or
normalized CPT tip resistance. Example charts used to estimate liquefaction induced settlement
using the Tokimatsu and Seed procedure and the Ishihara and Yoshimine procedure are presented
as Figure 6 -10 and Figure 6 -11, respectively.
Non-plastic to low plasticity silts (PI 12) have also been found to be susceptible to volumetric strain
following liquefaction. In cases where saturated silt is liquefiable the post-cyclic loading volumetric
strain should be estimated from cyclic laboratory testing, or approximately from the relationship
developed by Ishihara and Yoshimine.
Volume 1
47

ODOT Geotechnical Design Manual


April 2010

Figure 6-10. Liquefaction induced settlement estimated using the Tokimatsu & Seed
procedure (redrafted from Tokimatsu and Seed, 1987).

Volume 1
48

ODOT Geotechnical Design Manual


April 2010

Figure 6-11. Liquefaction induced settlement estimated using the Ishihara and Yoshimine
procedure. (redrafted from Ishihara and Yoshimine, 1992).

6.5.2.4

Residual Strength Parameters

Liquefaction induced instability is strongly influenced by the residual strength of the liquefied soil.
Instability occurs when the shear stresses required to maintain equilibrium exceed the residual
Volume 1
49

ODOT Geotechnical Design Manual


April 2010

strength of the soil deposit. Evaluation of residual strength of a liquefied soil deposit is one of the
most difficult problems in geotechnical practice. A variety of methods are available to estimate the
residual strength of liquefied soils. The procedures recommended in Section 6.4.1, include Seed and
Harder (1990), Olson and Stark (2002), Idriss (2003) and Kramer (2008). Other methods as
described in Dickenson, et. al., (2002), as well as the results of cyclic laboratory testing, may also be
used.
All of these methods estimate the residual strength of a liquefied soil deposit based on an empirical
relationship between residual undrained shear strength and equivalent clean sand SPT blowcounts
using the results of back-calculation of the apparent shear strengths from case histories, including
flow slides. All of these methods should be used to calculate the residual undrained shear strength
and an average value selected based on engineering judgment, taking into consideration the basis
and limitations of each correlation method.

6.5.3

Slope Stability and Deformation Analysis

Sloping earth structures and native slopes can become unstable due to: 1) liquefaction, or increased
pore pressures in soils, associated with a seismic event, 2) inertial effects associated with ground
accelerations, or 3) both. The methods described in this section, in Dickenson et. al (2002), and the
reference, should be used to assess seismic slope stability and for estimating ground displacements.
The slopes and conditions requiring such assessments and analysis are described in Section 6.2.4.
If liquefaction is not present, ground accelerations may produce inertial forces within the slope or
embankment that could exceed the strength of the foundation soils and result in slope failure and
large displacements. At these sites a pseudo-static analysis, which includes earthquake induced
inertia forces, is conducted to determine the general stability of the slope or embankment under
these conditions, as described in Section 6.5.3.1. The pseudo-static analysis is also used to
determine the yield acceleration which is then used in estimating slope or embankment
displacements.
If soils vulnerable to cyclic degradation (liquefiable soils, sensitive soils, brittle soils) are present slope
instability may develop in the form of flow failures, lateral spreading or other large embankment
deformations. Conventional slope stability analysis methods are typically conducted for liquefiable
soil sites using residual strength parameters to model the liquefied soils. The results of the analysis
are used to assess the potential for flow failures (FOS<1.0) and for use in displacement analysis.
Under liquefied soil conditions, slope stability is usually modeled in the post-earthquake condition
without including any inertial force from the earthquake ground motions (a de-coupled analysis) as
described in Section 6.5.3.2.

6.5.3.1

P s e u d o - s t a t i c An a l ys i s

Pseudo-static slope stability analyses should be used to evaluate the seismic stability of slopes and
embankments. The pseudo-static analysis shall consist of conventional limit equilibrium static slope
stability analysis, using horizontal and vertical pseudo-static acceleration coefficients (kh and kv) that
act upon the critical failure mass. Refer to Dickenson et al. (2002) for a discussion and detailed
guidance on pseudo-static analysis.
Non-liquefied soil conditions: For non-liquefied soil conditions, a horizontal pseudo-static
coefficient, kh, of 0.5As and a vertical pseudo-static coefficient, kv, equal to zero should be used when
seismic stability of slopes is evaluated. For these conditions, the minimum allowable factor of safety
is 1.0. Pseudo-static analyses do not result in predictions or estimates of slope deformation and
therefore are not sufficient for evaluation of bridge approach fill performance or for evaluating
foundations at the service limit state. The pseudo-static analysis is generally used to determine a
Volume 1
50

ODOT Geotechnical Design Manual


April 2010

yield acceleration for use in the Newmark (or other) analysis for estimating ground displacements, as
described in Section 6.5.3.2.
Liquefiable soil conditions: For liquefiable soil conditions, the potential for flow failures should be
assessed. Flow failures are driven by large static stresses that lead to large deformations or flow
following triggering of liquefaction. Such failures are similar to debris flows. Flow failures typically
occur near the end of strong shaking or shortly after shaking. However, delayed flow failures caused
by post-earthquake redistribution of pore water pressures can occurparticularly if liquefiable soils
are capped by relatively impermeable layers. For flow failures, both stability and deformation should
be assessed and mitigated if stability failure or excessive deformation is predicted.
Conventional limit equilibrium slope stability analysis methods are most often used to assess flow
failure potential. Residual undrained shear strength parameters are used to model the strength of the
liquefied soil. When using liquefied soil strengths, the horizontal and vertical pseudo-static
coefficients, kh and kv, respectively, should be set equal to zero (de-coupled analysis). Alternatively, a
site-specific, non-linear effective stress ground response analysis may be conducted to more
thoroughly assess liquefaction effects and determine appropriate acceleration values to use in the
pseudo-static analysis.
Where the factor of safety is less than 1.0, flow failure shall be considered likely. In these instances,
the magnitude of deformation is usually too large to be acceptable for design of bridges or structures,
and some form of mitigation may be appropriate. The exception is where the liquefied material and
crust flow past the structure and the structure can accommodate the imposed loads see
Section 6.5.6. Where the factor of safety is greater than 1.0, deformations can be estimated using
the methods described in Section 6.5.3.2.

6.5.3.2

D e f o r m a t i o n An a l ys i s

Deformation analyses should be employed where estimates of the magnitude of seismically induced
slope deformation are required. This is especially important for bridge approach fills where the
deformation analysis is a crucial step in evaluating whether or not the bridge performance
requirements described in Section 6.2 will be met. The procedures for estimating ground
deformations and examples are provided in Dickenson et al. (2002) and Dickenson (2005) along with
a discussion of which procedures are appropriate for specific conditions. It is recommended that
several of the methods described in these reports be used as appropriate and engineering judgment
applied to the results to determine the most reasonable range of predicted displacements.
Acceptable methods of estimating the magnitude of seismically induced lateral slope deformation
include:

Empirically-based displacement estimates for lateral spreading (Youd et al.

(2002),
Newmark-type analyses using acceleration time histories generated from sitespecific soil response modeling.

Simplified charts based on Newmark-type analyses (Makdisi and Seed, 1978)

Simplified procedures based on refined Newmark-type analyses (Bray and


Travasarou 2007, Saygili and Rathje 2008)

Simplified charts based on nonlinear, effective stress modeling (Dickenson et al,

2002)

Two-dimensional numerical modeling of dynamic slope deformation.

Volume 1
51

ODOT Geotechnical Design Manual


April 2010

The Newmark sliding block methods should not be employed to estimate displacements associated
with liquefaction or cyclic strength loss if the static factor of safety with the reduced (residual) strength
parameters is less than 1.0.
Brief summaries of each method are described below.
Youd et al. (2002); Lateral Spreading: If the slope stability factor of safety from the flow failure
analysis Section 6.5.3.1, assuming liquefied conditions, is 1.0 or greater, a lateral
spreading/deformation analysis should be conducted. Lateral spreading results when the shear
strength of the liquefied soil is incrementally exceeded by the inertial forces induced during an
earthquake. The result of lateral spreading is typically horizontal movement of nonliquefied soils
located above liquefied soils, in addition to the liquefied soils themselves.
The potential for liquefaction induced lateral spreading on gently sloping sites or where the site is
located near a free face may be evaluated using empirical relationships such as the procedure of
Youd et al. (2002). This procedure uses empirical relationships based on case histories of lateral
spreading. Input into the Youd et al. model includes earthquake magnitude, source-to-site distance,
and site geometry/slope, cumulative thickness of saturated soil layers and their characteristics (e.g.
SPT N values, average fines content and average grain size). This method is based on a
regression analysis of several independent variables correlated to field measurements of lateral
spread. Therefore it is best applied to site conditions that fit within the range of the variables used in
the models. Care should be taken when applying this method to sites with conditions outside the
range of the model variables. The Youd et al. procedure can provide a useful approximation of the
potential magnitude of deformation for sites with liquefiable soils.
Newmark Analysis: Newmark (1965) proposed a seismic slope stability analysis that provides an
estimate of seismically induced slope deformation. The advantage of the Newmark analysis over
pseudo-static analysis is that it provides an index of permanent deformation. The Newmark analysis
treats the unstable soil mass as a rigid block on an inclined plane. The procedure for the Newmark
analysis consists of three steps that can generally be described as follows:
Identify the yield acceleration of the slope by completing limit equilibrium stability
analyses. The yield acceleration is the horizontal pseudo-static coefficient, kh, required
to bring the factor of safety to unity (1.0). Note that if the yield acceleration applied to the
entire acceleration time history is based on residual soil strengths consistent with fully
liquefied conditions, the estimated lateral deformation will likely be overly conservative
since the liquefied, residual soil strength condition (and associated yield acceleration)
will only be in effect over a portion of the entire time history.

Select earthquake time histories representative of the design earthquakes. A


minimum of three time histories representative of the predominant earthquake source
zone(s) should be selected for this analysis. Note that these time histories need to be
propagated through the soil column to the ground surface to adjust them for local site
effects.

Double integrate all relative accelerations (i.e., the difference between


acceleration and yield acceleration) in the earthquake time histories.

A number of commercially available computer programs are available to complete Newmark


analysis, such as Shake2000 or Java Programs for using Newmarks Method and Simplified
Decoupled Analysis to Model Slope Performance during Earthquakes (Jibson, 2003).

Volume 1
52

ODOT Geotechnical Design Manual


April 2010

Makdisi-Seed Analysis: Makdisi and Seed (1978) developed a simplified procedure for estimating
seismically induced slope deformations based on Newmark sliding block analysis. The Makdisi-Seed
procedure provides an estimated range of permanent seismically induced slope deformation as a
function of the ratio of yield acceleration over maximum acceleration and earthquake magnitude as
shown on Figure 6 -12. The Makdisi-Seed procedure provides a useful index of the magnitude of
slope deformation. Because the Makdisi-Seed procedure includes the dynamic effects of the seismic
response of dams, its results should be interpreted with caution when applied to other slopes.

Figure 6-12. The Makdisi-Seed procedure for estimating the range of permanent
seismically induced slope deformation as a function of the ratio of yield acceleration
over maximum acceleration (redrafted from Makdisi and Seed, 1978).
Refined Newmark-Type Analysis-Bray and Travasarou (2007): This method is another
modification, or enhancement, of the original Newmark sliding block model. It consists of a simplified,
semiempirical approach for estimating permanent displacements due to earthquake-induced
deviatoric deformations using a nonlinear, fully coupled, stick-slip sliding block model. In addition to
estimating permanent displacements from rigid body slippage (basic Newmark approach) it also
includes estimates of permanent displacement (volumetric staining) from shearing within the sliding
mass itself. The model can be used to predict the probability of exceeding certain permanent
displacements or for estimating the displacement for a single deterministic event. This procedure is
available in EXCEL spreadsheet form.
Volume 1
53

ODOT Geotechnical Design Manual


April 2010

Refined Newmark-Type Analysis Saygili and Rathje (2008): This method is another
modification, or enhancement, of the original Newmark sliding block model, suitable for shallow
sliding surfaces that can be approximated by a rigid sliding block. The model predicts displacements
based on multiple ground motion parameters in an effort to reduce the standard deviation of the
predicted displacements.
Bracketed Intensity Method: The bracket intensity method is a modification of the Arias Intensity
procedure developed by Jibson (1993). The primary difference between the two methods is that the
bracketed intensity is a measure of only the ground motion intensity that is actually contributing to the
displacement of the sliding block. This method is quite similar to the Newmark-type methods. A stepby-step procedure for calculating the estimated displacement is presented in Dickenson et al. (2002).
Numerical Modeling Correlations (GMI): This is a simplified method for estimating lateral
deformations of embankments over liquefied soils. The method is presented Dickenson et al. (2002))
and is based on two dimensional numerical modeling of typical approach embankments using a finite
difference computer code (FLAC). In the procedure developed, limit equilibrium methods are used to
first calculate the post-earthquake factor of safety, using residual shear strengths in liquefied soils as
appropriate. The resulting FOS is then used in combination with a Ground Motion Intensity (GMI)
parameter to estimate embankment displacements. The GMI was developed to account for the
intensity and duration of the ground motions used in the FLAC analysis and is equal to the PGA
divided by the MSF (magnitude scaling factor). This procedure is also useful for estimating the
amount, or area, of ground improvement needed to limit displacements to acceptable levels.
Numerical Modeling of Dynamic Slope Deformation: Seismically induced slope deformations can
be estimated through a variety of dynamic stress-deformation computer models such as PLAXIS,
DYNAFLOW, and FLAC. The accuracy of these models is highly dependent upon the quality of the
input parameters. As the quality of the constitutive models used in dynamic stress-deformation
models improves, the accuracy of these methods will improve. Another benefit of these models is
their ability to illustrate mechanisms of deformation, which can provide useful insight into the proper
input for simplified analyses.
Dynamic stress deformation models should not be used for routine design due to their complexity,
and due to the sensitivity of the accuracy of deformation estimates from these models on the
constitutive model selected and the accuracy of the input parameters.

6.5.4

Settlement of Dry Sand

Seismically induced settlement of unsaturated granular soils (dry sands) is well documented. Factors
that affect the magnitude of settlement include the density and thickness of the soil deposit and the
magnitude of seismic loading. The most common means of estimating the magnitude of dry sand
settlement are through empirical relationships based on procedures similar to the Simplified
Procedure for evaluating liquefaction susceptibility. The procedures provided by Tokimatsu and Seed
(1987) for dry sand settlement should be used. The Tokimatsu and Seed approach estimates the
volumetric strain as a function of cyclic shear strain and relative density or normalized SPT N 60
values. The step-by-step procedure is presented in Section 8.5 of Geotechnical Engineering Circular
No. 3 (Kavazanjian, et al., 1997).

6.5.5

Liquefaction Effects on Structure Foundations

6.5.5.1

B r i d g e Ap p r o a c h F i l l s

All bridge approach fills should be assessed for the potential of excessive embankment deformation
(lateral displacement and settlement) due to seismic loading and the effects of these displacements
Volume 1
54

ODOT Geotechnical Design Manual


April 2010

on the stability and functional performance requirements of the bridge. This is true whether
liquefaction of the foundation soils is predicted or not. As a general rule, for the 500-year event, up to
one (1) foot of lateral and 6 to 12 inches of vertical embankment displacement can be used as a
general guideline for determining adequate performance of the approach fill. This range of
displacements should be considered only as a general guideline for evaluating the final condition of
the roadway surface and the ability to provide one-lane access to the bridge for emergency response
vehicles following the earthquake. Always keep in mind the accuracy of the methods used to predict
embankment deformations. Lateral displacement and fill settlement will also produce loads on the
bridge foundation elements which also have to be evaluated in terms of providing the required overall
bridge stability and performance. Specific embankment displacement limits are not provided for the
1000-year event since under this level of shaking the bridge and approach fills are evaluated only in
terms of meeting the No-Collapse criteria.

6.5.5.2

G e n e r a l L i q u e f a c t i o n P o l i c i e s Re g a r d i n g B r i d g e
Foundations

If liquefaction is predicted under either the 500 or 1000 year return events, the effects of liquefaction
on foundation design and performance must be evaluated. Soil liquefaction and the associated
effects of liquefaction on foundation resistances and stiffness is generally assumed, in standard
analyses, to be concurrent with the peak loads in the structure (i.e. no reduction in the transfer of
seismic energy due to liquefaction and soil softening). This applies except for the case where a sitespecific nonlinear effective stress ground response analysis is performed which takes into account
pore water pressure increases (liquefaction) and soil softening.
Liquefaction effects include:

reduced axial and lateral capacities and stiffness in deep foundations,

lateral spread,
embankments,

global

instabilities

and

displacements

of

slopes

and

ground settlement and possible downdrag effects

The following design practice, related to liquefied foundation conditions, should be followed:
Spread Footings: Spread footings are not recommended for bridge or abutment
wall foundation support over liquefiable soils unless ground improvement techniques are
employed that eliminate the potential liquefaction condition.

Piles and Drilled Shafts: The tips of piles and drilled shafts shall be located
below the deepest liquefiable soil layer. Friction resistance from liquefied soils should not
be included in either compression or uplift resistance recommendations for the Extreme
Event Limit I state loading condition. As stated above, liquefaction of foundation soils,
and the accompanying lose of soil strength, is assumed to be concurrent with the peak
loads in the structure. If applicable, reduced frictional resistance should also be applied
to partially liquefied soils either above or below the predicted liquefied layer. Methods for
this procedure are presented in Seed and Idriss, (1971) and Dickenson et al. (2002).

Pile Design Alternatives: Obtaining adequate lateral pile resistance is generally the main concern at
pier locations where liquefaction is predicted. Battered piles are not recommended. Prestressed
concrete piles have not been recommended in the past due to problems with excessive bending
stresses at the pile-footing connection. Vertical steel piles are generally recommended in high
seismic areas to provide the most flexible, ductile and cost-effective pile foundation system. Steel
pipe piles often are preferred over H-piles due to their uniform section properties, versatility in driving
with either closed or open ended and their potential for filling with reinforced concrete. The following
Volume 1
55

ODOT Geotechnical Design Manual


April 2010

design alternatives should be considered for increasing group resistance or stiffness and the most
economical design selected:

Increase pile size, wall thickness (section modulus) and/or strength.

Increase numbers of piles.

Increase pile spacing to reduce group efficiency effects.

Deepen pile cap and/or specify high quality backfill around pile cap for increase
capacity and stiffness.

Design pile cap embedment for fixed conditions.

Ground improvement techniques.

Liquefied P-y Curves: Studies have shown that liquefied soils retain a reduced, or residual, shear
strength and this shear strength may be used in evaluating the lateral capacity of foundation soils. In
light of the complexity of liquefied soils behavior (including progressive strength loss, strain
mobilization, and possible dilation and associated increase in soil stiffness) computer programs
commonly used for modeling lateral pile performance under liquefied soil conditions often rely on
simplified relationships for soil-pile interaction. At this time, no consensus exists within the
professional community on the preferred approach to modeling lateral pile response in liquefied soil.
Peer review is recommended for projects involving deep foundations in liquefiable soils.
One simplified procedure for modeling pile performance utilizes the static sand model(s) in the LPILE
program, modified using the residual shear strength and the effective overburden stress, at the depth
at which the residual strength was calculated (or measured), to estimate a reduced soil friction angle
(r) and initial soil modulus (k). The reduced soil friction angle is calculated using the inverse tangent
(i.e., Tan-1) of the residual undrained shear strength divided by the effective vertical stress at the
depth where the residual shear strength was determined or measured. Other procedures can be
used with approval by ODOT.
The reduced soil parameters may be applied to either the L-Pile static model or the strain wedge
static model (i.e., in DFSAP). DFSAP has an option built in to the program for estimating liquefied
lateral stiffness parameters and lateral spread loads on a single pile or shaft. However, it should be
noted the accuracy of the liquefied soil stiffness and predicted lateral spread loads using strain
wedge theory, in particular the DFSAP program, has not been well established and is not
recommended at this time,
For pile or shaft groups, for fully liquefied conditions, P-y curve group reduction factors may be set to
1.0. For partially liquefied conditions, the group reduction factors shall be consistent with the group
reduction factors used for static loading.
Additional liquefied P-y curve recommendations are provided in the research report titled: TILT: The
Treasure Island Liquefaction Test: Final Report, (Ashford, S. and Rollins, K., 2002) available from the
HQ Bridge Engineering Section. This full scale pile load test study produced P-y curves for liquefied
sand conditions that are fundamentally different than those derived from the standard static P-y curve
models. These liquefied P-y curves are available in Version 5 of the LPile computer program,
however the use of these liquefied p-y curves is not recommended at this time until further studies
are completed and a consensus is reached on the standard of practice for P-y curves to use in
modeling liquefied soils.
T-Z curves: Modify either the PL/AE method or APILE Plus program as follows:

Volume 1
56

ODOT Geotechnical Design Manual


April 2010

For the PL/AE method, if the liquefied zone reduces total pile skin friction to less
than 50% of ultimate bearing capacity, use end bearing condition (i.e. full length of pile)
in stiffness calculations. Otherwise use friction pile condition.

For the APile program, assume sand layer for liquefied zone with modified soil
input parameters similar to methods for P-y curve development.

Settlement and Downdrag Loads: Settlement of foundation soils due to the liquefaction or dynamic
densification of unsaturated cohesionless soils could result in downdrag loads on foundation piling or
shafts. Refer to Section 3.11.8 of the AASHTO LRFD Bridge Design Specifications for guidance on
designing for liquefaction-induced downdrag loads. Refer to Chapter 8 for guidance on including
seismic-induced settlement and downdrag loads on the seismic design of pile and shaft foundations.

6.5.5.3

L a t e r a l S p r e a d / S l o p e F a i l u r e L o ad s o n
Structures

In general, there are two different approaches to estimate the induced load on deep foundations
systems due lateral spreading a displacement based method and a force based method.
Displacement based methods are more prevalent in the United States. The force based approach
has been specified in the Japanese codes and is based on case histories from past earthquakes,
especially the pile foundation failures observed during the 1995 Kobe earthquake. Overviews of both
approaches are presented in the following sections.

6.5.5.4

D i s p l a c e m e n t B a s ed Ap p r o a c h

The recommended displacement based approach for evaluating the impact of liquefaction induced
lateral spreading loads on deep foundation systems is presented in the NCHRP Report 472 titled
Comprehensive Specification for Seismic Design of Bridges (NCHRP, 2002) and supporting
documentation by Martin et al., (2002). The general procedure is as follows:
Evaluate the Liquefaction Potential: Evaluate the liquefaction potential of the
site for the design risk levels (500 and 1000-yr. return periods). Assign residual and
reduced strength parameters to liquefied and partially liquefied soils layers.

Conduct Slope Stability Analyses: If liquefaction is predicted, conduct slope


stability analyses using residual strength parameters for the liquefied soil layers and
reduced strength parameters for partially liquefied soil layers. If the static factor of safety
is less than 1.0, a flow failure is predicted. If the static factor of safety is greater than 1.0,
conduct pseudo-static stability analyses to determine the yield acceleration Ky.

Check Zone of Influence: Assess whether or not the estimated failure surface
could impact the bridge foundation system. If the bridge foundations are expected to be
within the zone of influence, estimate the ground deformations.

Slope Deformations: Estimate lateral displacements based on the procedures


described in Section 6.5.3.2. Use all appropriate methods that could apply to the site
conditions and use judgment to determine the most reasonable amount of predicted
displacement that could occur.

Induced Loads on Foundation Elements: Assess whether the soil will displace
and flow around a stable foundation or whether foundation movement will occur in
concert with the soil. This assessment requires a comparison between the estimated
passive soil forces that can be exerted on the foundation and the ultimate resistance that
can be provided by the structure.

Volume 1
57

ODOT Geotechnical Design Manual


April 2010

The magnitudes of moment and shear induced in the foundations by the ground displacement can
be estimated using soil-pile structure interaction programs, such as LPile. The process is to apply the
assumed displacement field to the interface springs whose properties are represented by P-y curves.
The liquefied soil layers are typically modeled in the LPile or DFSAP programs using the modified
sand P-y model and the undrained residual strength of the liquefied soil see Section 6.5.5.2. Partially
liquefied soil layers are typically adjusted by reducing their friction angle (see Dickenson et al. 2003
for methods to reduce the friction angle based on increased pore pressures). The strength
parameters of non-liquefied layers above and/or below the liquefied zones are not reduced.
The estimated induced loads are then checked against the ability of the foundation system to resist
those loads. The ultimate foundation resistance is based in part on the resistance provided by the
portion of the pile/shaft embedded in non-liquefiable soils below the lateral spread zone and the
structural capacity of the pile/shaft. Large pile deformations may result in plastic hinges forming in the
pile/shaft. If foundation resistance is greater than that applied by the lateral spreading soil, the soil will
flow around the structure. If the potential load applied by the soil is greater than the ultimate
foundation system resistance, the pile/shaft is likely to move in concert with the soil. Also, the passive
pressure generated on the pile cap by the spreading soil needs to be considered in the total load
applied to the foundation system. In cases where a significant crust of non-liquefiable material may
exist, the foundation is likely to continue to move with the soil. Since large-scale structural
deformations may be difficult and costly to accommodate in design, mitigation of foundation subsoils
will likely be required.
In-ground hinging and plastic failure of piles or shafts due to lateral spread and slope failures is not
permitted on ODOT bridge projects for either the 500 or 1000 year design events.
Similar approaches to those outlined above can be used to estimate loads that other types of slope
failure may have on the bridge foundation system.

6.5.5.5

F o r c e B a s e d Ap p r o a c h e s

A force-based approach to assess lateral spreading induced loads on deep foundations is specified
in the Japanese codes. The method is based on back-calculations from pile foundation failures
caused by lateral spreading. The pressures on pile foundations are simply specified as follows:
The liquefied soil exerts a pressure equal to 30 percent of the total overburden
pressure (lateral earth pressure coefficient of 0.30 applied to the total vertical stress).

Non-liquefied crustal layers exert full passive pressure on the foundation system.

Data from simulated earthquake loading of model piles in liquefiable sands in centrifuge tests indicate
that the Japanese force method is an adequate design method (Finn, 2004).
Another force-based approach to estimate lateral spreading induced foundation loads is to use a limit
equilibrium slope stability program to determine the load the foundation must resist to achieve a
target safety factor of 1.1. This force is distributed over the foundation in the liquefiable zone as a
uniform stress. This approach may be utilized to estimate the forces that foundation elements must
withstand if they are to act as shear elements stabilizing the slope.

6.5.6

Mitigation Alternatives for Lateral Spread

The two basic options to mitigate the lateral spread induced loads on the foundation system are to
design the structure to accommodate the loads or improve the ground such that the hazard does not
occur.

Volume 1
58

ODOT Geotechnical Design Manual


April 2010

Structural Options (design to accommodate imposed loads): The general structural


approach to design for the hazard is outlined below.

Step 1: If the soil is expected to displace around the foundation element, the foundation
is designed to withstand the passive force exerted on the foundation by the flowing soil
and any overlying layers, or crust, of resistant soil. In this case, the maximum loads
determined from the P-y springs for large deflections are applied to the pile/shaft, and the
pile/shaft is evaluated using a soil structure interaction program similar to L-Pile. The
pile/shaft stiffness, strength, and embedment are adjusted until the desired structural
response to the loading is achieved.
Note that it is customary to evaluate the lateral spread/slope failure induced loads
independently from the inertial forces caused by the shaking forces (i.e. the shaking force
loads and the lateral spread loads are typically not assumed to act concurrently). In most
cases this is reasonable since peak vibration response is likely to occur in advance of
maximum ground displacement, and displacement induced maximum shear and
moments will generally occur at deeper depths than those from inertial loading.

Step 2: If the assessment indicates that movement of the foundation is likely to occur in
concert with the soil, then the structure is evaluated for the maximum expected ground
displacement. In this case the soil loads are generally not the maximum possible (loads
at large displacements), but instead some fraction thereof. Again the P-y data for the soils
in question are used to estimate the loading.

If the deformations determined in step 2 are beyond tolerable limits for structural design, the
options are to a) re-evaluate the deformations based on the pinning or doweling action that
foundations provide as they cross a potential failure plane (with consideration of the foundation
strength; or b) re-design the foundation system to accommodate the anticipated loads. Simplified
procedures for evaluating the available resistance to slope movements provided by the
foundation pinning action are presented in (NCHRP, 2002) and (Martin, et al., 2002) and require
knowledge of the plastic moment and location of plastic hinges in the foundation elements. This
information should be provided by the bridge designer or structural consultant. The concept of
considering a plastic mechanism or hinging in the piles/shafts is tantamount to accepting
foundation damage.
With input from the structural designer regarding pinning resistance provided by the foundation
system, recalculate the estimated displacement based on the revised resistance levels. If the
structures behavior is acceptable under the revised displacement estimate, the design for
liquefaction induced lateral spreading is complete. If the performance is not acceptable, then the
foundation system should be redesigned or ground improvement should be considered.
It is sometimes cost prohibitive to design the bridge foundation system to resist the loads
imposed by liquefaction induced lateral loads, especially if the depth of liquefaction extends more
than about 20 feet below the ground surface and if a non-liquefied crust is part of the failure
surface. Ground improvement to mitigate the liquefaction hazard is the likely alternative if it is not
practical to design the foundation system to accommodate the lateral loads.
Ground Improvement: The need for ground improvement techniques to mitigate liquefaction
effects depends, in part, upon the type and amount of anticipated damage to the structure and
approach fills due to the effects of liquefaction and embankment deformation (both horizontal and
vertical). The performance criteria described in Section 6.2 should be followed. Ground
Improvement methods are described in Elias et al. (2000) and Chapter 11. All ground

Volume 1
59

ODOT Geotechnical Design Manual


April 2010

improvement designs required to mitigate the effects of soil liquefaction shall be reviewed by the
HQ Bridge Section.
If, under the 500-year event, the estimated bridge damage is sufficient to render the bridge out of
service for one lane of emergency traffic then ground improvement measures should be
undertaken. If, under the 1000-year event, estimated bridge damage results in the possible
collapse of a portion or all of the structure then ground improvement is recommended. A flow
chart of the ODOT Liquefaction Mitigation Procedures is provided in Appendix 6-C.
Ground improvement techniques should result in reducing estimated ground and embankment
displacements to acceptable levels. Mitigation of liquefiable soils beneath approach fills should
extend a distance away, in both longitudinal and transverse directions, from the bridge abutment
sufficient enough to limit lateral embankment displacements to acceptable levels. As a general
rule of thumb, foundation mitigation should extend at least from the toe of the end slope to a point
where a 1:1 slope extending from the back of the bridge end panel intersects the original ground
(Figure 6 -13). The final limits of the mitigation area required should be determined from iterative
slope stability analysis and consideration of ground deformations. Practice-oriented procedures
have been described in Dickenson et al (2002).
Existing Grade

Bridge End Panel (typ. 30 ft.)


2:1 (typ.) Bridge
End Slope

1:1

Original Ground
Limits of Mitigation
Figure 6-13. Lateral Extent of Ground Improvement for Liquefaction Mitigation
Ground improvement techniques should also be considered as part of any Phase II (substructure
& foundation) seismic retrofit process. All Phase II retrofit structures should be evaluated for
liquefaction potential and mitigation needs. The cost of liquefaction mitigation for retrofitted
structures should be assessed relative to available funding.
The primary ground improvement techniques to mitigate liquefaction fall into three general
categories, namely densification, altering the soil composition, and enhanced drainage. A general
discussion regarding these ground improvement approaches is provided below.

Densification and Reinforcement: Ground improvement by densification consists of


sufficiently compacting the soil such that it is no longer susceptible to liquefaction during a
design seismic event. Densification techniques include vibro-compaction, vibro-flotation,
vibro-replacement (stone columns), deep dynamic compaction, blasting, and compaction
grouting. Vibro-replacement and compaction grouting also reinforce the soil by creating
columns of stone and grout, respectively. The primary parameters for selection include
grain size distribution of the soils being improved, depth to groundwater, depth of
improvement required, proximity to settlement/vibration sensitive infrastructure, and
access constraints.

Altering Soil Composition: Altering the composition of the soil typically refers to
changing the soil matrix so that it is no longer susceptible to liquefaction. Examples of

Volume 1
60

ODOT Geotechnical Design Manual


April 2010

ground improvement techniques include permeation grouting (either chemical or microfine cement), jet grouting, and deep soil mixing. These types of ground improvement are
typically more costly than the densification/reinforcement techniques, but may be the
most effective techniques if access is limited, construction induced vibrations must be
kept to a minimum, and/or the improved ground has secondary functions, such as a
seepage barrier or shoring wall.

Drainage Enhancements: By improving the drainage properties of sandy soils


susceptible to liquefaction, it may be possible to reduce the build-up of excess pore water
pressures, and thus liquefaction during seismic loading. However, drainage improvement
is not considered adequately reliable by ODOT to prevent excess pore water pressure
buildup due to the length of the drainage path, the time for pore pressure to dissipate, the
influence of fines on the permeability of the sand, and due to the potential for drainage
structures to become clogged during installation and in service. In addition, with drainage
enhancements some settlement is still likely. Therefore, drainage enhancements alone
shall not be used as a means to mitigate liquefaction.

Geotechnical engineers are encouraged to work with ground treatment contractors having
regional experience in the development of soil improvement strategies for mitigating hazards due
to permanent ground deformation.

Input for Structural Design

6.6.1

Foundation Springs

Structural dynamic response analyses incorporate the foundation stiffness into the dynamic model of
the structure to capture the effects of soil structure interaction. The foundation stiffness is typically
represented as a system of equivalent springs placed in a foundation stiffness matrix. The typical
foundation stiffness matrix incorporates a set of six springs, namely a vertical spring, horizontal
springs in the orthogonal plan dimensions, rocking about each horizontal axis, and torsion around the
vertical axis.
The primary parameters for calculating the individual springs are the foundation type (shallow spread
footings or deep foundations), foundation geometry, and dynamic soil shear modulus. The dynamic
soil shear modulus is a function of the shear strain (foundation displacement), so determining the
appropriate foundation springs can be an iterative process. Refer to the ODOT BDDM for additional
information on foundation modeling methods and the soil/rock design parameters required by the
structural designer for the analysis.

6.6.1.1

S h a l l o w F o u n d at i o n s

For evaluating shallow foundation springs, the structure designer requires values for the dynamic
shear modulus, G, Poissons ratio, and the unit weight of the foundation soils. The maximum, or lowstrain, shear modulus can be estimated using index properties and the correlations presented in
Table 6 -14. Alternatively, the maximum shear modulus can be calculated using Equation 6.2, if the
shear wave velocity is known:
Gmax = /g(Vs)2
Where:
Gmax = maximum dynamic shear modulus
= soil unit weight
Volume 1
61

ODOT Geotechnical Design Manual


April 2010

Vs = shear wave velocity


g = acceleration due to gravity
The maximum dynamic shear modulus is associated with small shear strains (less than 0.0001
percent). As shear strain level increases, dynamic shear modulus decreases. At large cyclic shear
strain (1 percent), the dynamic shear modulus approaches a value of approximately 10 percent of
Gmax (Seed et al., 1986). As a minimum, shear modulus values for 0.2 percent shear strain and 0.02
percent shear strain to simulate large and small magnitude earthquakes should be provided to the
structural engineer. Shear modulus values at other shear strains could also be provided as needed
for the design. Shear modulus values may be estimated using Figure 6 -2, Figure 6 -3 and Figure 6
-4. Alternatively, laboratory tests, such as the cyclic triaxial or direct simple shear, or resonant column
tests may be used to determine the shear modulus values at intermediate shear strains. The results
of in situ tests such as the CPT and DMT have also been used to develop non-linear relationships for
soil stiffness (Mayne, 2001).
Poissons Ratio can be estimated based on soil type, relative density/consistency of the soils, and
correlation charts such as those presented in Foundation Analysis and Design (Bowles, 1996).

6.6.1.2

Deep Foundations

Lateral soil springs for deep foundations shall be determined in accordance with Chapter 8. Refer to
Section 6.5.5.2 for guidance on modifying t-z curves and the soil input required for P-y curves
representing liquefied or partially liquefied soils.

6.6.1.3

D o w n d ra g L o ad s o n S t r u c t u r e s

Downdrag loads on foundations shall be determined in accordance with Chapter 8.

Volume 1
62

ODOT Geotechnical Design Manual


April 2010

References

AASHTO, 2007. AASHTO LRFD Bridge Design Specifications, Fourth Edition.


AASHTO, 2009. AASHTO Guide Specifications for LRFD Seismic Bridge Design.
AASHTO, 1988, Manual on Subsurface Investigations.
American Association of State Highway and Transportation Officials (AASHTO). AASHTO LRFD
Bridge Design Specifications, Customary U.S. Units. 5th Edition, with 2010 Interim Revisions.
AASHTO, 2010.
Atwater, Brian F., 1996. Coastal Evidence for Great Earthquakes in Western Washington. Assessing
Earthquake Hazards and Reducing Risk in the Pacific Northwest, USGS Professional Paper 1560
Vol. 1: pp. 77-90.
Ashford, S. and Rollins, K., 2002. TILT: The Treasure Island Liquefaction Test: Final Report,
Department of Structural Engineering, Univ. of California, San Diego, Report No. SSRP-2001/17.
Bakun, W.H., Haugerud, R.A., Hopper, M.G., and Ludwin, R.S., 2002. The December 1872
Washington State Earthquake, Bulletin of the Seismological Society of America, Vol. 92, No. 8, pp.
3239-3258.
Boulanger, R. W. and Idriss, I. M., 2006. Liquefaction Susceptibility Criteria for Silts and Clays,
ASCE Journal of Geotechnical and Geoenvironmental Engineering, Vol. 132, No. 11, pp. 1413-1426.
Bowles, J.E., 1996. Foundation Analysis and Design, Fifth Edition. The McGraw-Hill Companies,
Inc., New York.
Bray, J. and Rathje, E., 1998. Earthquake Induced Displacements of Solid Waste Landfills, ASCE
Journal of Geotechnical and Geoenvironmental Engineering, Vol. 124, pp. 242-253.
Bray, J. D., and Sancio, R. B., 2006. Assessment of the Liquefaction Susceptibility of Fine-Grained
Soil, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 132, No. 9, pp. 11651176.
Bray, J. D., and Travasarou, T, 2007. Simplified Procedure for Estimating Earthquake-Induced
Deviatoric Slope Displacements, Journal of Geotechnical and Geoenvironmental Engineering,
ASCE, Vol. 133, No. 4, pp. 381-392.
Volume 1
63

ODOT Geotechnical Design Manual


April 2010

Dickenson, S., McCullough, N., Barkau, M. and Wavra, B., 2002. Assessment and Mitigation of
Liquefaction Hazards to Bridge Approach Embankments in Oregon, ODOT Final Report SPR
361.
Dickenson, S., 2005. Recommended Guidelines for Liquefaction Evaluations Using Ground
Motions from Probabilistic Seismic Hazard Analysis, Report to ODOT.
Report 02-420, U.S. DEPARTMENT OF THE INTERIOR U.S. GEOLOGICAL SURVEY, 2002,
Documentation for the 2002 Update of the National Seismic Hazard Maps", Open-File.
Electrical Power Research Institute (EPRI), 1993. Guidelines for Site Specific Ground Motions. Palo
Alto, CA. Electrical Power Research Institute, November-TR-102293.
Elias, V., Welsh, J., Warren, J., and Lukas, R., 2000, Ground Improvement Technical Summaries
Vol. 1 and 2, Demonstration Project 116, Federal Highway Administration, FHWA-SA-98-086.
Finn, W.D. Liam, Ledbetter, R.H. and Wu, G., 1994. Liquefaction in Silty Soils: Design and Analysis.
Ground Failures Under Seismic Conditions, Geotechnical Special Publication 44. ASCE, New York,
New York, pp. 51-76.
Finn, W.D. Liam and Fujita, N., 2004. Behavior of Piles in Liquefiable Soils during Earthquakes:
Analysis and Design Issues. Proceedings: Fifth International Conference on Case Histories in
Geotechnical Engineering, New York, New York, April 13-17, 2004.
Goter, S.K., 1994. Earthquakes in Washington and Oregon; 1872-1993, 1994, USGS Open-File
Report No. 94-226A.
Idriss, I. M., 2003, Resolved and unresolved issues in soil liquefaction, Presentation at US-Taiwan
Workshop on Soil Liquefaction Hsinchu, Taiwan, November 2.
Induced Ground Failure Hazards, Transportation Research Board. National Research Council,
Washington, D.C.
International Code Council, Inc., 2002. 2003 International Building Code. Country Club Hills, IL.
Ishihara, K., and Yoshimine, M., 1992. Evaluation of settlements in sand deposits following
liquefaction during earthquakes, Soils and Foundations, JSSMFE, Vol. 32, No. 1, March, pp. 173188.

Volume 1
64

ODOT Geotechnical Design Manual


April 2010

Jibson R.W., (1993). Predicting Earthquake Induced Landslide Displacements Using Newmarks
Sliding Block Analysis, Transportation Research Record, No. 1411-Earthqauke Induced Ground
Failure Hazards. Transportation Research Board. National Research Council, Washington, D.C.
Jibson R. and Jibson M., 2003. Java Program for using Newmarks Method and Simplified
Decoupled Analysis to Model Slope Deformations During Earthquakes. Computer Software. USGS
Open File Report 03-005.
Kavazanjian, E., Matasovic, N., Hadj-Hamou, T. and Sabatini, P.J., 1997. Geotechnical Engineering
Circular #3, Design Guidance: Geotechnical Earthquake Engineering for Highways, Volume I: Design
Principles and Volume II: Design Examples. Report Nos. FHWA-SA-97-076/077. U.S. Department of
Transportation, Federal Highway Administration, Washington, D.C.
Kramer, S.L., 1996. Geotechnical Earthquake Engineering. Prentice-Hall, Inc., Upper Saddle River,
NJ.
Kramer, S.L. and Paulsen, S.B., 2004. Practical Use of Geotechnical Site Response Models. PEER
Lifelines Program Workshop on the Uncertainties in Nonlinear Soil Properties and the Impact on
Modeling Dynamic Soil Response. Berkeley, CA. March 18-19, 2004.
Lee, M. and Finn, W., 1978. DESRA-2, Dynamic Effective Stress Response Analysis of Soil Deposits
with Energy Transmitting Boundary Including Assessment of Liquefaction Potential. Soil Mechanics
Series No. 38, Dept. of Civil Engineering, University of British Columbia, Vancouver, B.C.
Makdisi, F.I. and Seed, H.B., 1978. Simplified Procedure for Estimating Dam and Embankment
Earthquake-Induced Deformations. ASCE Journal of the Geotechnical Engineering Division, Vol.
104, No. GT7, July, 1978, pp. 849-867.
Martin, G.R., Marsh, M.L., Anderson, D.G., Mayes, R.L., and Power, M.S., 2002. Recommended
Design Approach for Liquefaction Induced Lateral Spreads. Third National Seismic Conference and
Workshop on Bridges and Highways, Portland, Oregon, April 28 May 1, 2002.
Matasovic, N. and Vucetic, M. (1995b), Seismic Response of Soil Deposits Composed of FullySaturated Clay and Sand, Proceedings of the 1st International Conference on Geotechnical
Earthquake Engineering, Tokyo, Japan, Vol. I, pp. 611-616.
Mayne, P. W., 2001. Proceedings, International Conference on In-Situ Measurement of Soil
Properties & Case Histories [In-Situ 2001]; Bali, Indonesia, May 21-24, pp. 27-48.
McGuire, R.K., 2004. Seismic Hazard and Risk Analysis. Monograph MNO-10, Earthquake
Engineering Research Institute, Oakland, CA. 221 pp.
Volume 1
65

ODOT Geotechnical Design Manual


April 2010

National Cooperative Highway Research Program, 2002. Comprehensive Specification for the
Seismic Design of Bridges, NCHRP Report 472, Washington DC. Newmark, N.M., 1965. Effects of
Earthquakes on Dams and Embankments. Geotechnique 15(2), pp. 139-160.
Oregon Department of Transportation, 2005, Bridge Design and Drafting Manual.
Ordoez, G.A., 2000. Shake 2000, Computer Software.
Sabatini, P.J., Bachus, R.C., Mayne, P.W., Schneider, J.A., and Zettler, T.E., 2002. Geotechnical
Engineering Circular No. 5, Evaluation of Soil and Rock Properties, Report No. FHWA-IF-02034. U.S. Department of Transportation, Federal Highway Administration, Washington, D.C.
Satake, Kenji, et al., 1996. Time and Size of a Giant Earthquake in Cascadia Inferred from
Japanese Tsunami Records of January 1700. Nature, Vol. 379, pp. 247-248.
Saygili, G., and Rathje, E. M., 2008, Empirical Predictive Models for Earthquake-Induced Sliding
Displacements of Slopes, ASCE Journal of Geotechnical and Geoenvironmental Engineering, Vol.
134, No. 6, pp. 790-803.
Seed, H.B. and Idriss, I.M., 1970. Soil Moduli and Damping Factors for Dynamic Response Analysis.
Report No. EERC 70-10, University of California, Berkeley.
Seed, H.B. and Idriss, I.M., 1971. Simplified Procedure for Evaluating Soil Liquefaction Potential.
ASCE Journal of Soil Mechanics and Foundations Division, Vol. 97, No. SM9, pp. 1249-1273.
Seed, H.B., Wong, R. T., Idriss, I.M., and Tokimatsu, K., 1986. Moduli and Damping Factors for
Dynamic Analyses of Cohesionless Soils. ASCE Journal of Geotechnical Engineering, Vol. 112, No.
11, pp. 1016-1032.
Seed, R.B. and Harder, L.F., 1990. SPT-based Analysis of Cyclic Pore Pressure Generation and
Undrained Residual Strength. Proceedings, H. Bolton Seed Memorial Symposium, University of
California, Berkeley, Vol. 2, pp. 351-376.
Stewart, J.P., Liu, A.H. and Choi, Y., 2003. Amplification Factors for Spectral Acceleration in
Tectonically Active Regions. Bulletin of Seismological Society of America, Vol. 93 No. n1 pp. 332352.
Tokimatsu, K. and Seed, H.B., 1987. Evaluation of Settlement in Sands Due to Earthquake
Shaking. ASCE Journal of Geotechnical Engineering, Vol. 113, No. 8, August 1987.
United States Geological Survey, 2002. Earthquake Hazards Program. Website link:
Volume 1
66

ODOT Geotechnical Design Manual


April 2010

http://earthquake.usgs.gov/hazmaps/
Vucetic, M. and Dobry, R. (1991). Effect of Soil Plasticity on Cyclic Response. Journal of
Geotechnical Engineering, Vo. 117, No. 1, pp. 89-107.
Yelin, T.S., Tarr, A.C., Michael, J.A., and Weaver, C.S., 1994. Washington and Oregon Earthquake
History and Hazards. USGS, Open File Report 94-226B.
Youd, T.L.; Idriss, I.M.; Andrus, R.D.; Arango, I.; Castro, G.; Christian, J.T.; Dobry, R.; Finn, W.D.;
Harder, L.; Hynes, M.E.; Ishihara, K.; Koester, J.P.; Liao, S.S.C.; Marcuson, W.F.; Martin, G.R.;
Mitchell, J.K.; Moriwaki, Y.; Power, M.S.; Robertson, P.K.; Seed, R.B. and Stokoe, K.H., 2001.
Liquefaction Resistance of Soils: Summary Report from the 1996 NCEER and 1998 NCEER/NSF
Workshops on Evaluation of Liquefaction Resistance of Soils. ASCE Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 127, No. 10, pp. 817-833.
Youd, T.L.; Hansen, C.M. and Bartlett, S.F., 2002. Revised Multilinear Regression Equations for
Prediction of Lateral Spread Displacement. ASCE Journal of Geotechnical and Geoenvironmental
Engineering, Vol. 128, No. 12, pp. 1007-1017.

Volume 1
67

ODOT Geotechnical Design Manual


April 2010

Appendix 6-A
FLOW CHART FOR EVALUATION OF LIQUEFACTION HAZARD
AND GROUND DEFORMATION AT BRIDGE SITES

STEP 1
Identify Seismic Sources in the Region
CSZ interplate, deep intraplate, shallow crustal earthquakes refer to USGS Seismic Hazard Mapping Project Web Site
Obtain M-R pairs from de-aggregation tables for 475 and 975 mean return periods
Consider the following sources:
CSZ Interplate Earthquakes
M 8.3 and M 9.0
as defined by the USGS

Deep Intraplate Earthquake


Very small contribution to PGA
hazard in most of Oregon
Confirm on De-Aggregation tables
by checking for representative M-R
pairs

Crustal, Areal, or Gridded Seismicity


Obtain M-R pairs from USGS deaggregation tables for all regional
Define criteria for selecting all M-R pairs
that significantly contribute to the overall
seismic hazard

STEP 2
Select Appropriate Ground Motion Attenuation Relationships for each Source and Style of Faulting
Calculate the bedrock PGA values for each M-R pair
STEP 3
Select Appropriate Acceleration Time Histories for Bedrock Motions
Three, or more, records from different earthquakes are recommended per M-R pair
Consider style of faulting, magnitude, and the characteristics of the candidate motions (duration,
frequency content, and energy)

STEP 4
Perform Dynamic Soil Response Analysis
Develop profiles of cyclic stress ratio (CSR) versus depth for each M-R pair (3 or more time histories per M-R pair)
Compute the average CSR profile with depth for each M-R pair
Compute suite of Acceleration Response Spectra (ARS) if needed for structural engineering

STEP 5
Compute the Factor of Safety against Liquefaction for each M-R Pair
Use the averaged CSR profile for each M-R pair
Utilize standard methods for liquefaction susceptibility evaluation based on penetration resistance or shear wave velocity

Volume 1
68

ODOT Geotechnical Design Manual


April 2010

STEP 6
Establish Post-Cyclic Loading Shear Strengths of Embankment and Foundation
Soils

This is performed for each M-R pair


Focus on sensitive soils, weak fine-grained soils, loose to medium dense sandy soils (potentially
liquefiable soils are addressed as follows)

If FSliq 1.4

If FSliq 1.0

If 1.4 > FSliq > 1.0

Use drained shear


strengths

Estimate the
residual excess pore
pressure
Compute the

Estimate the residual undrained


strength using two or more
methods

equivalent friction angle


STEP 7
Perform Slope Stability Analysis

Static analysis using post-cyclic loading shear strengths for each M-R pair
Calculate the FOS against sliding and determine the critical acceleration values for each M-R pair
Focus trial slip surfaces on weak soil layers

STEP 8
Perform Deformation Analysis for each M-R pair

Rigid-body, sliding block analysis (Newmark Method)


Simplified chart solutions
Numerical modeling

STEP 9
Evaluate Computed Deformations in Terms of Tolerable Limits

Permanent Deformations are Unacceptable


Permanent Deformations are
Acceptable
Computed displacements are
less than defined limits
Continue with structural
design

Computed displacements exceed defined limits repeat


analysis incorporating the effects of remedial ground treatment

Return to Step 4 if the soil improvement does not


significantly change the anticipated dynamic response of the soil
column (e.g., isolated soil improvement)

Return to Step 3 if the ground treatment substantially


alters the dynamic response of the site (e.g., extensive soil
improvement in the vertical and lateral direction, extensive
treatment including grouting or deep soil mixing)

A reduced number of input time histories are acceptable


for each
M-R pair (bracket the problem using trends from the initial analysis)

Volume 1
69

ODOT Geotechnical Design Manual


April 2010

Appendix 6-B: Example of Smoothed Response Spectra

Volume 1
70

ODOT Geotechnical Design Manual


April 2010

Chapter 6 D5

4/20/2015

Appendix 6-C: ODOT Liquefaction Mitigation


Procedures
Foundation Design Engineer evaluates liquefaction potential using the
500 yr. event and estimates approach fill deformations
(Lateral displacement, settlement and global stability)
No
Check liquefaction and
est. displacements under
1000 yr. event

No

Is there potential for large embankment


deformations? (see Note 1 below)
Yes

Geotechnical and Structural Designers meet and determine damage potential


to structure and serviceability of bridge. Will the bridge and/or approaches be
damaged such that the bridge will be out of service?
(see Note 2 below)

Geotechnical and
Structural Designers
determine damage
potential to structure and
possibility of collapse

Typical Design

Yes

No
Is there a
possibility of
bridge collapse?
Yes

Proceed with Mitigation Design


Alternatives (Note 3)

Note 1: For meeting the performance requirements of the 500 year return event (serviceability), lateral deformation of
approach fills of up to 12 are generally considered acceptable under most circumstances pending an evaluation of this
amount of lateral deformation on abutment piling. Larger lateral deformations and settlements may be acceptable under
the 1000 year event as long as the no-collapse criteria are met.
Note 2: The bridge should be open to emergency vehicles after the 500-year design event, following a thorough
inspection. If the estimated embankment deformations (vertical or horizontal or both) are sufficient enough to cause
concerns regarding the serviceability of the bridge mitigation is recommended.
Note 3: Refer to ODOT research report SPR Project 361: Assessment and Mitigation of Liquefaction Hazards to Bridge
Approach Embankments in Oregon, Nov. 2002 and FHWA Demonstration Project 116; Ground Improvement Technical
Summaries, Volumes I & II, (Pub. No. FHWA-SA-98-086) for mitigation alternatives and design procedures.
As a general guideline, the foundation mitigation should extend from the toe of the end slope to a point that is located at
the base of a 1:1 slope which starts at the end of the bridge end panel:

Existing Grade

Bridge End Panel (typ. 30 ft.)


2:1 (typ.)
1:1

Original Ground
Limits of Mitigation

Volume 1
71

ODOT Geotechnical Design Manual


April 2010

Chapter 6 D1

4/20/2015

Volume 1
72

ODOT Geotechnical Design Manual


April 2010

Chapter

Slope Stability Analysis

General

Slope stability analysis is used in a wide variety of geotechnical engineering problems, including, but
not limited to, the following:

Determination of stable cut and fill slopes

Assessment of overall stability of retaining walls, including global and compound stability
(includes permanent systems and temporary shoring systems)

Assessment of overall stability of shallow and deep foundations for structures located on
slopes or over potentially unstable soils, including the determination of lateral forces applied
to foundations and walls due to potentially unstable slopes

Stability assessment of landslides (mechanisms of failure, and determination of design


properties through back-analysis), and design of mitigation techniques to improve stability

Evaluation of instability due liquefaction

Types of slope stability analyses include rotational slope failure, sliding block analysis, irregular
surfaces of sliding, and infinite slope failure. Stability analysis techniques specific to rock slopes,
other than highly fractured rock masses, that can in effect be treated as soil, are described in
Chapter 12. Detailed stability assessment of landslides is described in Chapter 13.

Development of
and Input Data
Analysis

Design Parameters
for Slope Stability

The input data needed for slope stability analysis is described in Chapter 2 for site investigations in
general, Chapter 9, Chapter 10 for fills and cuts, and Chapter 13 for landslides. Chapter 5
provides requirements for the assessment of design property input parameters. Detailed assessment
of soil and rock stratigraphy is critical to the proper assessment of slope stability, and is in itself a
direct input parameter for slope stability analysis. It is important to define any thin weak layers
Volume 1
1

ODOT Geotechnical Design Manual


April 2010

present, the presence of slickensides, etc., as these fine details of the stratigraphy could control the
stability of the slope in question. Knowledge of the geologic nature of the units present at the site and
knowledge of past performance of such units may also be critical factors in the assessment of slope
stability. Whether long-term or short-term stability is in view, and which will control the stability of the
slope, will affect the selection of soil and rock shear strength parameters used as input in the
analysis. For short-term stability analysis, undrained shear strength parameters should be obtained.
For long-term stability analysis, drained shear strength parameters should be obtained. For
assessing the stability of landslides, residual shear strength parameters will be needed, since the soil
has in such cases already deformed enough to reach a residual value. For highly over-consolidated
clays, if the slope is relatively free to deform after the cut is made or is otherwise unloaded, residual
shear strength parameters should be obtained and used for the stability analysis.
Detailed assessment of the groundwater regime within and beneath the slope is also critical. Detailed
piezometric data at multiple locations and depths within and below the slope will likely be needed,
depending on the geologic complexity of the stratigraphy and groundwater conditions. Potential
seepage at the face of the slope must be assessed and addressed. In some cases, detailed flow net
analysis may be needed. If seepage does exit the slope face, the potential for soil piping should also
be assessed as a slope stability failure mechanism, especially in highly erodable silts and sands.

Design Requirements

Limit equilibrium methodologies are usually used to assess slope stability. The Modified Bishop,
simplified Janbu, Spencer, or other widely accepted slope stability analysis methods should be used
for rotational and irregular surface failure mechanisms. In cases where the stability failure
mechanisms anticipated are not well modeled by limit equilibrium techniques, or if deformation
analysis of the slope is required, more sophisticated analysis techniques (e.g., finite difference
methodologies such as is used by the computer program FLAC) may be used in addition to the limit
equilibrium methodologies. Since these more sophisticated methodologies are quite sensitive to the
quality of the input data and the details of the model setup, including the selection of constitutive
models used to represent the material properties and behavior, limit equilibrium methods should also
be used in such cases. If the differences in the results are significant, engineering judgment should
be applied in conjunction with any available field observations to assess the correctness of the design
model used. If the potential slope failure mechanism is anticipated to be relatively shallow and
parallel to the slope face, with or without seepage affects, an infinite slope analysis should be
conducted. Typically, slope heights of 15 to 20 ft or more are required to have this type of failure
mechanism. For infinite slopes which are either above the water table or which are fully submerged,
the factor of safety for slope stability is determined as follows:

Seepage: Considering that the buoyant unit weight is roughly one-half of the saturated unit
weight, seepage on the slope face can reduce the factor of safety by a factor of two, a
condition which should obviously be avoided through some type of drainage if at all possible;
otherwise much flatter slopes will be needed.

Slopes: When using the infinite slope method, if the FS is near or below 1.0 to 1.15, severe
erosion or shallow slumping is likely. Vegetation on the slope can help to reduce this problem,
as the vegetation roots add cohesion to the surficial soil, improving stability. Note that
conducting an infinite slope analysis does not preclude the need to check for deeper slope
failure mechanisms, such as would be assessed by the Modified Bishop or similar methods
listed above. For very simplified cases, design charts to assess slope stability are available.
Examples of simplified design charts are provided in NAVFAC DM-7. These charts are for a
c- soil, and apply only to relatively uniform soil conditions within and below the cut slope.
They do not apply to fills over relatively soft ground, as well as to cuts in primarily cohesive

Volume 1
2

ODOT Geotechnical Design Manual


April 2010

soils. Since these charts are for a c- soil, a small cohesion will be needed to perform the
calculation.
If these charts are to be used, it is recommended that a cohesion value of 50 to 100 psf be
used in combination with the soil friction angle obtained from SPT correlation for relatively
clean sands and gravels.

Soil parameters: For silty to very silty sands and gravels, the cohesion could be increased to
100 to 200 psf, but with the friction angle from SPT correlation (see Chapter 5) reduced by 2
to 3 degrees, if it is not feasible to obtain undisturbed soil samples suitable for laboratory
testing to measure the soil shear strength directly. This should be considered general
guidance, and good engineering judgment should be applied when selecting soil parameters
for this type of an analysis. Simplified design charts should only be used for final design of
non-critical slopes that are 10 ft in height or less and that are consistent the simplified
assumptions used by the design chart. Simplified design charts may be used as applicable
for larger slopes for preliminary design. The detailed guidance for slope stability analysis
provided by Abramson, et al. (1996) should be used.

Resistance Factors and Safety Factors


for Slope Stability Analysis

For overall stability analysis of walls and structure foundations, design shall be consistent with
Chapter 6, Chapter 8 and Chapter 15 and the AASHTO LRFD Bridge Design Specifications. For
slopes adjacent to but not directly supporting structures, a maximum resistance factor of 0.75 should
be used. For foundations on slopes that support structures such as bridges and retaining walls, a
maximum resistance factor of 0.65 should be used. Exceptions to this could include minor walls that
have a minimal impact on the stability of the existing slope, in which the 0.75 resistance factor may
be used. Since these resistance factors are combined with a load factor of 1.0 (overall stability is
assessed as a service limit state only), these resistance factors of 0.75 and 0.65 are equivalent to a
safety factor of 1.3 and 1.5, respectively. For general slope stability analysis of cuts, fills, and
landslide repairs, a minimum safety factor of 1.25 should be used. Larger safety factors should be
used if there is significant uncertainty in the slope analysis input parameters. For seismic analysis, if
seismic analysis is conducted see Chapter 6 for policies on this issue, a maximum resistance factor
of 0.9 should be used for slopes involving or adjacent to walls and structure foundations. This is
equivalent to a safety factor of 1.1. For other slopes (cuts, fills, and landslide repairs), a minimum
safety factor of 1.05 should be used.

References

Abramson, L., Boyce, G., Lee, T., and Sharma, S., 1996, Slope Stability and Stabilization Methods,
Wiley, ISBN 0471106224.

Volume 1
3

ODOT Geotechnical Design Manual


April 2010

Chapter

Foundation Design

General

This chapter covers the geotechnical design of bridge foundations, retaining wall foundations and
cut-and-cover tunnel foundations. Both shallow and deep foundation types are addressed.
Foundation design work entails assembling all available foundation information for a structure,
obtaining additional information as required, performing foundation analyses and compiling the
information into a report that includes the specific structure foundation recommendations. An
adequate site inspection, office study, appropriate subsurface exploration program and
comprehensive foundation analyses that result in foundation recommendations are all necessary to
construct a safe, cost-effective structure. See Chapter 21 for guidance on the foundation information
that should be included in Geotechnical Reports. See Chapter 2 for guidance on foundation
information available through office studies and the procedures for conducting a thorough site
reconnaissance.
Unless otherwise stated in this manual, the Load and Resistance Factor Design approach (LRFD)
shall be used for all foundation design projects, as prescribed in the most current version of the
AASHTO LRFD Bridge Design Specifications. The ODOT foundation design policies and standards
described in this chapter supersede those in the AASHTO LRFD specifications. FHWA design
manuals are also acceptable for use in foundation design and preferable in cases where foundation
design guidance is not adequately provided in AASHTO. Structural design of bridge foundations, and
other structure foundations, is addressed in the ODOT Bridge Design and Drafting Manual (BDDM).
It is important to establish and maintain close communication between the geotechnical designer and
the structural designer at all times throughout the entire foundation design process and continuing
through construction.

Project Data and Foundation Design


Requirements

The scope of the project, project requirements, project constraints and the geology and subsurface
conditions of the site should be analyzed to determine the type and quantity of geotechnical
Volume 1
1

ODOT Geotechnical Design Manual


April 2010

investigation work to be performed. Project information such as a vicinity map, a project narrative,
preliminary structure plans/layout (pre-Type, Size & Location) and hydraulics information (if
applicable) should be obtained to allow for proper planning of the subsurface exploration program.
Keep abreast of changes to the project scope that might impact the geotechnical investigation and
design work required. Proposed retaining wall and bridge bent locations should be obtained from the
bridge designer prior to the beginning of field work to properly locate bore holes.
Anticipated foundation loads, structure settlement criteria and the heights of any proposed fills should
be determined or estimated to insure that the exploration boreholes are advanced to the proper
depth and the proper information is obtained.
Refer to AASHTO Article 10.4.1 for more details of the information needed at this stage.
The foundation type(s) selected for each structure will each require specific subsurface investigation
methods, materials testing, analysis and design. Table 8 -15 provides a summary of information
needs and testing considerations for foundation design.
Table 8-15. Summary of information needs and testing considerations (modified after Sabatini, et. al.
2002)
Foundation
Type
Shallow
Foundations

Engineering
Evaluations

bearing capacity
settlement
(magnitude &
rate)
shrink/swell of
foundation soils
(natural soils or
embankment fill)
frost heave
scour (for water
crossings)
liquefaction

Required
Information
For Analyses
subsurface profile (soil,
groundwater, rock)
shear strength parameters
compressibility parameters
(including consolidation,
shrink/swell potential, and
elastic modulus)
frost depth
stress history (present and
past vertical effective
stresses)
depth of seasonal moisture
change
unit weights
geologic mapping including
orientation and
characteristics of rock
discontinuities

Volume 1
2

Field Testing

SPT (granular
soils)
CPT
PMT
dilatometer
rock coring
(RQD)
plate load
testing
geophysical
testing

Laboratory
Testing

1-D Oedometer
tests
soil/rock shear
tests
grain size
distribution
Atterberg Limits
specific gravity
moisture content
unit weight
organic content
collapse/ swell
potential tests
intact rock modulus
point load strength
test

ODOT Geotechnical Design Manual


April 2010

Table 8-1 (Continued)


Driven Pile
Foundations

Drilled Shaft
Foundations

pile end-bearing
pile skin friction
settlement
down-drag on pile
lateral earth
pressures
chemical
compatibility of
soil and pile
drivability
presence of
boulders/very
hard layers
scour (for water
crossings)
vibration/heave
damage to nearby
structures
liquefaction
shaft end bearing
shaft skin friction
constructability
down-drag on
shaft
quality of rock
socket
lateral earth
pressures
settlement
(magnitude &
rate)
groundwater
seepage/
deepwatering/
potential for
caving
presence of
boulders/very
hard layers
scour (for water
crossings)
liquefaction

subsurface profile (soil,


groundwater, rock)
shear strength parameters
horizontal earth pressure
coefficients
interface friction
parameters (soil and pile)
compressibility parameters
chemical composition of
soil/rock (e.g., potential
corrosion issues)
unit weights
presence of shrink/swell
soils (limits skin friction)
geologic mapping including
orientation and
characteristics of rock
discontinuities

subsurface profile (soil,


groundwater, rock)
shear strength parameters
interface shear strength
friction parameters (soil
and shaft)
compressibility parameters
horizontal earth pressure
coefficients
chemical composition of
soil/rock
unit weights
permeability of waterbearing soils
presence of artesian
conditions
presence of shrink/swell
soils (limits skin friction)
geologic mapping including
orientation and
characteristics of rock
discontinuities
degradation of soft rock in
presence of water and/or
air (e.g., rock sockets in
shales)

Volume 1
3

SPT (granular
soils)
pile load test
CPT
PMT
vane shear
test
dilatometer
piezometers
rock coring
(RQD)
geophysical
testing

installation
technique test
shaft
shaft load test
vane shear
test
CPT
SPT (granular
soils)
PMT
dilatometer
piezometers
rock coring
(RQD)
geophysical
testing

soil/rock shear
tests
interface friction
tests
grain size
distribution
1-D Oedometer
tests
pH, resistivity tests
Atterberg Limits
specific gravity
organic content
collapse/ swell
potential tests
intact rock modulus
point load strength
test

1-D Oedometer
tests
soil/rock shear
tests
grain size
distribution
interface friction
tests
pH, resistivity tests
permeability tests
Atterberg Limits
specific gravity
moisture content
unit weight
organic content
collapse/ swell
potential tests
intact rock modulus
point load strength
test
slake durability

ODOT Geotechnical Design Manual


April 2010

Field Exploration for Foundations

Subsurface explorations shall be performed in accordance with Article 10.4.2 of the AASHTO LRFD
Bridge Design Specifications, supplemented by the FHWA Geotechnical Engineering Circular No. 5,
Evaluation of Soil and Rock Properties (FHWA-IF-02-034). The procedures outlined in the ODOT
Soil and Rock Classification Manual are used to describe and classify subsurface materials. The
explorations shall provide the information needed for the design and construction of foundations.
Accurate and adequate subsurface information at, or as near as possible to, each structure support is
extremely important, especially for drilled shaft and spread footing designs.
The minimum exploration requirements specified in AASHTO Section 10, and as supplemented in
Chapter 3, should be considered the standard of practice with regards to subsurface investigation
requirements. It is understood that engineering judgment will need to be applied by a licensed and
experienced geotechnical professional to adapt the exploration program to the foundation types and
depths needed and to the variability in the subsurface conditions observed. The extent of exploration
shall be based on the variability in the subsurface conditions, structure type, foundation loads, and
any project requirements that may affect the foundation design or construction. The exploration
program should be extensive enough to reveal the nature and types of soil deposits and/or rock
formations encountered the engineering properties of the soils and/or rocks, the potential for
liquefaction, and the groundwater conditions. The exploration program should be sufficient to identify
and delineate problematic subsurface conditions such as deep, very soft soil deposits, bouldery
deposits, swelling or collapsing soils, existing fill or waste areas, etc.
For cut-and-cover tunnels, culverts, arch pipes, etc., spacing of exploration locations shall be
consistent with the requirements described in Chapter 3.
The groundwater conditions at the site are very important for both the design and construction of
foundations. Groundwater conditions are especially important in the construction of drilled shafts,
spread footings or any other excavation that might extend below the water table or otherwise
encounter groundwater. Piezometer data adequate to define the limits and piezometric head in all
unconfined, confined, and locally perched groundwater zones should be obtained at each foundation
location. The measured depth and elevations of groundwater levels, and dates measured, should be
noted on the exploration logs and discussed in the final Geotechnical Report. It is important to
distinguish between the groundwater level and the level of any drilling fluid. Also, groundwater levels
encountered during exploration may differ from design groundwater levels. Any artesian
groundwater condition or other unusual groundwater condition should be identified and reported as
this often has important impacts on foundation design and construction.

Field and Laboratory


Foundations

Testing

for

Conduct subsurface investigations and materials testing in conformance with AASHTO Articles
10.4.3 and 10.4.4. Table 8 -15 provides a summary of field and laboratory testing considerations for
foundation design. Foundation design will typically rely upon the Standard Penetration Test (SPT),
Cone Penetrometer Test (CPT) and rock core samples obtained during the field exploration. Visual
descriptions of the soil and rock materials are recorded. Correlations are usually made between
these field tests to shear strength and compressibility of the soil. Groundwater and other hydraulic
information needed for foundation design and constructability evaluation is typically obtained during
the exploration using field instrumentation (e.g., piezometers) and in-situ tests (e.g., slug tests, pump
tests, etc.).
ODOT owns the following equipment:
Volume 1
4

ODOT Geotechnical Design Manual


April 2010

A Texam Pressuremeter which is available for use on Agency designed projects.


The pressuremeter requires predrilled boreholes. The pressuremeter is stored in Region
2. Contact the Region 2 Bridge/Geo-Hydro Section for assistance in obtaining the use of
this equipment.

A Vane Shear device, a Point Load Tester and a Geoprobe. Contact the
Pavements Unit to schedule use of the Geoprobe equipment.

In general, for foundation design, laboratory testing should be used to augment the data obtained
from the field investigation program and to refine the soil and rock properties selected for design.
Index tests such as soil gradation, Atterberg limits, water content, and organic content are used to
confirm the visual field classification of the soils encountered, but may also be used directly to obtain
input parameters for some aspects of foundation design (e.g., soil liquefaction, scour, degree of overconsolidation, and correlation to shear strength or compressibility of cohesive soils). Laboratory tests
conducted on undisturbed soil samples are used to assess shear strength or compressibility of finer
grained soils, or to obtain seismic design input parameters such as shear modulus.

Material Properties for Design

The selection of soil and rock design properties should be in conformance with those described in
Chapter 5 with additional reference to "Evaluation of Soil and Rock Properties, Geotechnical
Engineering Circular No. 5, (FHWA-IF-02-034).

Bridge Approach Embankments

The embankments at bridge ends should be evaluated for stability and settlement. The FHWA
publication Soils and Foundations Reference Manual, 2006, (FHWA NHI-06-088) should be
referenced for guidance in the analysis and design of bridge approach embankments. New
embankment placed for bridge approaches should be evaluated for short term (undrained) and long
term (drained) conditions.
Bridge end slopes are typically designed at 2(H):1(V). If steeper end slopes such as 1: 1 are
desired, they should be evaluated for stability and designed to meet the required factors of safety. If
embankment stability concerns arise, consider the use of staged construction, wick drains, flatter
slopes, soil reinforcement, lightweight materials, subexcavation/replacement, counterbalances, or
other measures depending on site conditions, costs and constraints. The embankment stability
analysis, any recommended stabilization measures, instrumentation or other embankment
monitoring needs, should be described in detail in the Geotechnical Report.
For overall stability, the static factor of safety for bridge approach embankments should be at least
1.30. A factor of safety of at least 1.5 must be provided against overall stability for abutment spread
footings supported directly on embankments or abutment retaining walls. The programs XSTABL5.2
and Slope/W are available for evaluating slope stability. Dynamic (seismic) slope stability, settlement
and lateral displacements are discussed in Chapter 6.
The FHWA program EMBANK (Urzua, A., 1993) is available for use in estimating embankment
settlement. If the estimated post-construction settlement is excessive, consider the use of waiting
periods, surcharges, wick drains or other ground improvement methods to expedite or minimize
embankment settlement and allow for bridge construction. Consider relocating the bridge end if
embankment settlement and stability concerns result in extreme and costly measures to facilitate
embankment construction. Also, evaluate long term embankment settlement potential and possible
downdrag effects on piles or drilled shafts and provide downdrag mitigation recommendations, such
as wait periods, if necessary. In general, design for the long term settlement of approach
Volume 1
5

ODOT Geotechnical Design Manual


April 2010

embankments to not exceed 1 in 20 years. Refer to the BDDM for additional approach fill
settlement limitations regarding integral abutments.

8.6.1

Abutment Transitions

ODOT standard practice is to provide bridge end panels at each end bent location for bridges
constructed on the State Highway system. Embankment settlement often occurs at this transition
point after construction is completed and the end panels are necessary to eliminate a potentially
dangerous traffic hazard and reduce the impact of traffic loads to the bridge. The settlement is
sometimes the result of poorly placed and compacted embankment material or abutment backfill it or
might be due to long-term settlement of the foundation soils. Guidance for proper detailing and
material requirements for abutment backfill is provided in the "Soils and Foundations Reference
Manual, 2006, (FHWA NHI-06-088).
End panels may be considered for deletion if the following geotechnical conditions are met::
Foundation materials are characterized as incompressible (e.g., bedrock or
very dense granular soils)

Post-construction settlement estimates are negligible (<0.25),

Provisions are made to insure the specifications for embankment and backfill
materials, placement and compaction are adhered to (increased inspection and testing
QC/QA)

A geotechnical and structural evaluation is required for considering the deletion of end panels and
approval of a deviation from standard ODOT BDDM practice is required. The final decision on
whether or not to delete end panels shall be made by the ODOT HQ Bridge Section Engineer with
consideration to the geotechnical and structural evaluation.
In addition to geotechnical criteria, other issues such as average daily traffic (ADT), design speed, or
accommodation of certain bridge structure details may supersede the geotechnical reasons for
deleting end panels. End panels shall be used for all ODOT bridges with stub, or integral abutments
to accommodate bridge expansion and contraction. End panels shall also be used in all cases where
seismic loads could result in excessive dynamic fill settlement and the failure to meet the
performance criteria described in the BDDM.

8.6.2

Overall Stability

The evaluation of overall stability of earth slopes with or without a foundation unit shall be
investigated at the service limit state as specified in Article 11.6.2.3 of the AASHTO LRFD Bridge
Design Specifications. Overall stability should be evaluated using limiting equilibrium methods such
as modified Bishop, Janbu, Spencer, or other widely accepted slope stability analysis methods.
Article 11.6.2.3 recommends that overall stability be evaluated at the Service I limit state (i.e., a load
factor of 1.0) and a resistance factor, os, of 0.65 for slopes which support a structural element. This
corresponds to a factor of safety of 1.5.
Most slope stability programs produce a single factor of safety, FS. Overall slope stability shall be
checked to insure that foundations designed for a maximum bearing stress equal to the specified
service limit state bearing resistance will not cause the slope stability factor of safety to fall below 1.5.
This practice will essentially produce the same result as specified in Article 11.6.2.3 of the AASHTO
LRFD Bridge Design Specifications. The foundation loads should be as specified for the Service I
limit state for this analysis. If the foundation is located on the slope such that the foundation load
contributes to slope instability, the designer shall establish a maximum footing load that is acceptable
for maintaining overall slope stability for Service, and Extreme Event limit states (see Figure 8 -14 for
Volume 1
6

ODOT Geotechnical Design Manual


April 2010

example). If the foundation is located on the lower portion of the slope such that the foundation load
increases slope stability, overall stability of the slope shall be evaluated ignoring the effect of the
footing on slope stability.

Figure 8-14. Example where footing contributes to instability of slope (left figure) vs.
example where footing contributes to stability of slope (right figure)

Foundation Selection Criteria

The foundation type selected for a given structure should result in the design of the most economical
bridge, taking into account any constructability issues and constraints. The selection of the most
suitable foundation for the structure should be based on the following considerations:
the ability of the foundation type to meet performance requirements (e.g.,
deformation, bearing resistance, uplift resistance, lateral resistance/deformation) for all
limit states including scour and seismic conditions,

the constructability of the foundation type (taking into account issues like traffic
staging requirements, construction access, shoring required, cofferdams)

the cost of the foundation,

meeting the requirements of environmental permits (e.g. in-water work periods,


confinement requirements, noise or vibration effects from pile driving or other operations,
hazardous materials)

constraints that may impact the foundation installation (e.g., overhead clearance,
access, surface obstructions, and utilities)

the impact of foundation construction on adjacent structures, or utilities, and the


post-construction impacts on such facilities,

the impact of the foundation installation (in terms of time and space required) on
traffic and right-of-way

This is the most important step in the foundation design process. These considerations should be
discussed (as applicable) with the structural designer. Bridge bent locations may need to be adjusted
Volume 1
7

ODOT Geotechnical Design Manual


April 2010

based on the foundation conditions, construction access or other factors described above to arrive at
the most economical and appropriate design.

Spread Footings
Spread footings are typically very cost effective, given the right set of conditions. Footings work best
in hard or dense soils that have adequate bearing resistance and exhibit tolerable settlement under
load. Footings can get rather large in less dense soils such as medium dense sand or stiff clays
depending on the structure loads and settlement requirements. Structures with tall columns or with
high lateral loads which result in large eccentricities and footing uplift loads may not be suitable
candidates for footing designs. Footings are not allowed or cost-effective where soil liquefaction can
occur at or below the footing level. Other factors that affect the cost feasibility of spread footings
include:

the need for a cofferdam and seals when placed below the water table,

the need for significant over-excavation and replacement of unsuitable soils,

the need to place footings deep due to scour, liquefaction or other conditions,

the need for significant shoring to protect adjacent existing facilities, and

inadequate overall stability when placed on slopes that have marginally adequate
stability.

Settlement (service limit state criteria) often controls the feasibility of spread footings. The amount of
footing settlement must be compatible with the overall bridge design. The superstructure type and
span lengths usually dictate the amount of settlement the structure can tolerate and footings may still
be feasible and cost effective if the structure can be designed to tolerate the estimated settlement
(e.g., flat slab bridges, bridges with jackable abutments, etc.). Footings may not be feasible where
expansive or collapsible soils are present near the bearing elevation. Refer to the FHWA
Geotechnical Engineering Circular No. 6, Shallow Foundations, and the FHWA publication, Selection
of Spread Footings on Soils to Support Highway Bridge Structures (FHWA-RC/TD-10-001) for
additional guidance on the selection and use of spread footings.

Deep Foundations
Deep foundations are the next choice when spread footings cannot be founded on competent soils or
rock at a reasonable cost. Deep foundations are also required at locations where footings are
unfeasible due to extensive scour depths, liquefaction or lateral spread problems. Deep foundations
may be installed to depths below these susceptible soils to provide adequate foundation resistance
and protection against these problems. Deep foundations should also be used where an
unacceptable amount of spread footing settlement may occur. Deep foundations should be used
where right-of-way, space limitations, or other constraints as discussed above would not allow the
use of spread footings.
Two general types of deep foundations are typically considered: pile foundations, and drilled shaft
foundations. The most economical deep foundation alternative should be selected unless there are
other controlling factors. Shaft foundations are most advantageous where very dense intermediate
strata must be penetrated to obtain the desired bearing, uplift, or lateral resistance, or where
materials such as boulders or logs must be penetrated. Shafts may also become cost effective where
a single shaft per column can be used in lieu of a pile group with a pile cap, especially when a
cofferdam, seal and/or shoring is required to construct the pile cap. However, shafts may not be
desirable where contaminated soils are present, since the contaminated soil removed would require
special handling and disposal. Constructability is also an important consideration in the selection of
Volume 1
8

ODOT Geotechnical Design Manual


April 2010

drilled shafts. Shafts can be used in lieu of piles where pile driving vibrations could cause damage to
existing adjacent facilities.
Piles may be more cost effective than shafts where pile cap construction is relatively easy, where the
depth to the foundation layer is large (e.g., more than 100 ft), or where the pier loads are such that
multiple shafts per column, requiring a shaft cap, are needed. The tendency of the upper loose soils
to flow, requiring permanent shaft casing, may also be a consideration that could make pile
foundations more cost effective. Artesian pressure in the bearing layer could preclude the use of
drilled shafts due to the difficulty in keeping enough head inside the shaft during excavation to
prevent heave or caving under slurry.
When designing pile foundations keep in mind the potential cost impacts associated with the use of
large pile hammers. Local pile driving contractors own hammers typically ranging up to about 80,000
ft.-lbs of energy. When larger hammers are required to drive piles to higher pile bearing resistance
they have to rent the hammers and the mobilization cost associated with furnishing pile driving
equipment may increase sharply. Larger hammers may also impact the design and cost work bridges
due to higher hammer and crane loads.
For situations where existing substructures must be retrofitted to improve foundation resistance,
where there is limited headroom available for pile driving or shaft construction, or where large
amounts of boulders must be penetrated, micropiles may be the best foundation alternative, and
should be considered.
Augercast piles can be very cost effective in certain situations. However, their ability to resist lateral
loads is minimal, making them undesirable to support structures where significant lateral loads must
be transferred to the foundations. Furthermore, quality assurance of augercast pile integrity and
capacity needs further development. Therefore, it is ODOT current policy not to use augercast piles
for bridge foundations.

Overview of LRFD for Foundations

The basic equation for load and resistance factor design (LRFD) states that the loads multiplied by
factors to account for uncertainty, ductility, importance, and redundancy must be less than or equal to
the available resistance multiplied by factors to account for variability and uncertainty in the
resistance per the AASHTO LRFD Bridge Design Specifications. The basic equation, therefore, is as
follows:
i Qi Rn

(8-1)

= Factor for ductility, redundancy, and importance of structure


i = Load factor applicable to the ith load Qi
Qi = Load
= Resistance factor
Rn = Nominal (predicted) resistance
For typical ODOT practice, is set equal to 1.0 for use of both minimum and maximum load factors.
The product, Rn, is termed the factored resistance. This term is analogous to the term allowable
capacity previously used in Allowable Stress Design. AASHTO Article 10.5.5 provides the resistance
factors to use in foundation design. Resistance factors for a given foundation type are a function of
the design method used, soil type/condition and other factors. AASHTO Article 10.5.5, and its
Volume 1
9

ODOT Geotechnical Design Manual


April 2010

associated commentary, should be reviewed for information on the development of the specified
resistance factors used in foundation design and provides guidance in the selection and use of these
factors. Foundations shall be proportioned so that the factored resistance is always greater than or
equal to the factored loads. The loads and load factors to be used in pile foundation design shall be
as specified in Section 3 of the AASHTO LRFD Bridge Design Specifications.

Foundation Design Policies

8.9.1

Downdrag Loads

Downdrag loads on piles, shafts, or other deep foundations shall be evaluated as described in
AASHTO Article 3.11.8. If a downdrag condition exists, the resulting downdrag loads (DD) are
included with the permanent load combinations used in structure design and an appropriate load
factor is applied to the downdrag loads. In addition to applying the downdrag loads on the load side
of the LRFD equation, the downdrag loads must also be subtracted from the resistance side of the
equation since this resistance will not be available for foundation support.
If the settlement cannot be mitigated, consideration should be given to reducing the effects of
downdrag loads on the foundations by the use of bitumen coating or pile sleeves. The NCHRP
Report Design and Construction Guidelines for Downdrag on Uncoated and Bitumen-Coated Piles
(Briaud, J. et al., 1997) should be referenced for more guidance on downdrag mitigation methods.
Earthquakes may also produce foundation settlement and downdrag loads due to either liquefaction
of saturated sandy soils or dynamic compaction of unsaturated sandy soils resulting from seismic
ground motions. Chapter 6 presents methods for calculating liquefaction potential and dynamic
settlement estimates. Downdrag loads resulting from seismic loading conditions should not be
combined with downdrag loads resulting from static long-term foundation settlement.

8.9.2

Scour Design

Structures crossing waterways may be subject to damage by scour and erosion of the streambed,
stream banks, and possibly the structure approach fills. Bents placed in the streambed increase the
potential for scour to occur. The degree and depth of scour will have a significant affect on the
selection of the most appropriate foundation type. The Hydraulic Report should be consulted for
scour predictions.
Scour depths are typically calculated for both the 100-year (base flood) and 500-year (check flood)
events. However, if overtopping of the roadway can occur, the incipient roadway overtopping
condition is then the worst case for scour because it will usually create the greatest flow contraction
and highest water velocities at the bridge. This overtopping condition may occur less than every 100
years and therefore over-ride the base flood (100-yr) design condition or it could occur between 100
and 500 years and over-ride the 500-year (check flood) condition. All bridge scour depths are
calculated for the following flood conditions, depending on the recurrence interval for the overtopping
flood:

Qovertopping > Q500: Both the 100-year and 500-year flood scour depths are analyzed

Q100 < Qovertopping < Q500 : The 100-year flood and the overtopping flood scour
depths are analyzed

Qovertopping < Q100: Only the overtopping flood scour depth is analyzed

The top of the footing should be set below the potential scour elevation for the 100-year scour or the
roadway-overtopping flood, whichever is the deepest. The bottom of the footing should be set below
Volume 1
10

ODOT Geotechnical Design Manual


April 2010

the potential scour elevation for the Check Flood, which will be either the roadway-overtopping flood
or the 500-year flood.
Minimum pile and drilled shaft tip elevations and spread footing elevations should be based on
providing the nominal bearing resistance (resistance factor equal to 1.0) with the estimated 500-year
flood scour depths or with the scour depths from the overtopping flood if the recurrence interval of the
overtopping flood is greater than 100-years. A resistance factor of 0.70 may be used in foundation
design with the estimated 100-year flood scour depths. However, if the recurrence interval of the
overtopping flood is less than 100 years, the resistance factor should be evaluated on a case by case
basis using engineering judgment and assessing the long term hydraulics and scour potential of the
site. Overtopping recurrence intervals that are much less than every 100 years are not considered
extreme events and therefore resistance factors associated with the no-scour condition may be more
appropriate to use.
For footings constructed on bedrock, provide recommendations regarding the scour potential of the
bedrock to the Hydraulics designer. Some types of bedrock are very weak and extremely
susceptible to erosion and scour. At present, there are no specific recommendations or guidelines to
use to determine the scour potential of bedrock types typically found in Oregon. Good engineering
judgment should be used in estimating the scour potential of marginally good quality rock, taking
into account rock strength, RQD, joint spacing, joint filling material, open fractures, weathering,
degradation characteristics and other factors. See if any exposed bedrock at the site shows signs of
erosion or degradation or if there is a history of bedrock scour in the past. Signs of bedrock scour
may include the undermining of existing footings, steeply incised stream banks or scour holes in the
bedrock streambed. If any doubts remain, drilled shafts should be considered.
Spread footings supporting bridge abutments should generally be constructed assuming the
contraction and degradation scour depths calculated for the main channel are present at the
abutment location. Exceptions to this policy include bridge abutment footings that are constructed on
non-erodable rock and/or located sufficiently far away from the main channel (e.g. long approach
ramps or viaduct). Refer to the ODOT Hydraulics Manual for more guidance regarding scour, riprap
protection and footing depth requirements. Loose riprap is not considered permanent protection.
Design riprap protected abutments according to the guidance and recommendations outlined in
FHWA HEC-18 manual, Evaluating Scour at Bridges (Richardson, E. et al., 2001).

8.9.3

Seismic Design

Chapter 6 describes ODOT seismic practices regarding design criteria, performance requirements,
ground motion characterization, liquefaction analysis, ground deformation and mitigation. Once the
seismic analysis is performed the results are applied to foundation design in the Extreme Event I limit
state analysis as described in AASHTO Section 10. Also refer to, and be familiar with, Section 1.1.4;
Foundation Modeling, of the ODOT Bridge Design and Drafting Manual.
This section describes the various methods bridge designers use to model the response of bridge
foundations to seismic loading and also the geotechnical information required to perform the
analysis.
In general, nominal resistances are used in seismic design except for pile and shaft uplift conditions
(see AASHTO Article 10.5.5.3).
If the foundation soils are determined to be susceptible to liquefaction, then spread footings should
not be recommended for foundation support of the structure unless proven ground improvement
techniques are employed to stabilize the foundation soils and eliminate the liquefaction potential.
Otherwise, a deep foundation should be recommended.

Volume 1
11

ODOT Geotechnical Design Manual


April 2010

Deep foundations (piles and drilled shafts) supporting structures that are constructed on potentially
liquefiable soils are normally structurally checked for two separate loading conditions; i.e. with and
without liquefaction. Nominal (unfactored) resistances, downdrag loads (if applicable) and soil-pile
interaction parameters should be provided for both nonliquefied and liquefied foundation conditions.
Communication with the structural designer is necessary to insure that the proper foundation design
information is provided.
If the predicted amount of earthquake-induced embankment deformation (lateral deformation and/or
settlement) is excessive then assessments should be made of approach fill performance and the
potential for bridge and approach fill damage. The need for possible liquefaction mitigation measures
should then be evaluated. Refer to the ODOT Liquefaction Mitigation Policy, in Chapter 6, for more
guidance on ODOT liquefaction mitigation policies.

Soil Loads on Buried Structures

For tunnels, culverts and pipe arches, the soil loads to be used for design shall be as specified in
Sections 3 and 12 of the AASHTO LRFD Bridge Design Specifications.

Spread Footing Design

Refer to AASHTO Article 10.6 for spread footing design requirements.


Once footings are selected as the preferred design alternative, the general spread footing design
process can be summarized as follows. Close communication and interaction is required between
the structural and geotechnical designers throughout the footing design phase.
Determine footing elevation based on location of suitable bearing stratum and
footing dimensions (taking into account any scour requirements, if applicable)

Determine foundation material design parameters and groundwater conditions

Calculate the nominal bearing resistance for various footing dimensions (consult
with structural designer for suitable dimensions)

Select resistance factors depending on design method(s) used; apply them to


calculated nominal resistances to determine factored resistances

Determine nominal bearing resistance at the service limit state

Check overall stability (determine max. bearing load that maintains adequate
slope stability)

For footings located in waterways, the bottom of the footing should be below the estimated depth of
scour for the check flood (typically the 500 year flood event or the overtopping flood). The top of the
footing should be below the depth of scour estimated for the design flood (either the overtopping or
100-year event). As a minimum, the bottom of all spread footings should also be at least 6 feet below
the lowest streambed elevation unless they are keyed full depth into bedrock that is judged not to
erode over the life of the structure. Spread footings are not permitted on soils that are predicted to
liquefy under the design seismic event.

8.11.1

Nearby Structures

Refer to AASHTO, Article 10.7.1.6.4. Issues to be investigated include, but are not limit to, settlement
of the existing structure due to the stress increase caused by the new footing, decreased overall
stability due to the additional load created by the new footing, and the effect on the existing structure
of excavation, shoring, and/or dewatering to construct the new foundation.
Volume 1
12

ODOT Geotechnical Design Manual


April 2010

8.11.2

Service Limit State Design of Footings

Footing foundations shall be designed at the service limit state to meet the tolerable movements for
the structure in accordance with AASHTO Article 10.5.2. The nominal unit bearing resistance at the
service limit state, qserve, shall be equal to or less than the maximum bearing stress that results in
settlement that meets the tolerable movement criteria for the structure.

Driven Pile Foundation Design

Refer to AASHTO, Article 10.7 for pile design requirements. Pile design should meet or exceed the
requirements specified for each limit state. ODOT standards and policies regarding pile foundation
design and construction shall also be followed.
All driven piles shall be accepted based on bearing resistance determined from dynamic formula,
wave equation, dynamic measurements with signal matching (PDA/CAPWAP) or full-scale load
testing. Pile acceptance shall not be accepted based solely on static analysis.
For piles requiring relatively low nominal resistances (<600 kips) and without concerns about high
driving stresses, the dynamic formula is typically used for determining pile driving acceptance criteria.
In cases where piles are driven to higher resistances or where high pile driving stresses are a
concern, such as short, end bearing piles, the wave equation is typically used for both drivability
analysis and in determining the final driving acceptance criteria.
Pile acceptance based on the pile driving analyzer (PDA) is typically reserved for projects where it is
economically advantageous to use, or for cases where high pile driving stresses are predicted and
require monitoring. The PDA (with signal matching) method can be most cost effective on projects
that have a large number of long, high capacity, friction piles.
Full-scale static pile load tests are less common in practice due to their inherent expense. However,
they may be economically justified in cases where higher bearing resistances can be verified through
load testing and applied in design to reduce the cost of the pile foundation. If static load testing is
considered for a project it should be conducted early on in the design stage so the results may be
utilized in the design of the structure. Also, the pile load test should be taken to complete failure if at
all possible. Refer to AASHTO Section 10 for descriptions on how to use the results of the static load
tests results to determine driving criteria. Static load test results should be used in combination with
either PDA testing or wave equation analysis to develop final driving criteria for the production piles.
Once the pile (bent) locations and foundation materials and properties are defined, the pile
foundation design process for normal bridge projects typically consists of the following:

Determine scour depths (if applicable)

Determine liquefaction potential and depths; estimate seismic induced settlement


(if applicable)

Evaluate long-term embankment settlement and downdrag potential

Select most appropriate pile type

Select pile dimension (size) based on discussions with structural designer


regarding preliminary pile loading requirements (axial and lateral)

Establish structural nominal resistance of the selected pile(s)

Conduct static analysis to calculate nominal single pile resistance as a function of


depth for the strength and extreme limit states (or a pile length for a specified resistance)

Volume 1
13

ODOT Geotechnical Design Manual


April 2010

Select resistance factors based on the field method to be used for pile
acceptance (e.g. dynamic formula, wave equation, PDA/CAPWAP, etc.)

Calculate single pile factored resistance as a function of depth

Estimate downdrag loads; consolidation and/or seismic-induced (if applicable)

Calculate pile/pile group settlement or pile lengths required to preclude excessive


settlement

Determine nominal uplift resistance as a function of depth

Determine p-y curve parameters for lateral load analysis


o

Modify parameters for liquefied soils (if applicable)

o
Provide p-y multipliers as appropriate for pile groups. P-Y multipliers are
not required for pile (or shafts) groups installed in rock sockets where calculated
lateral displacements are minimal (i.e., <0.50).
Determine required pile tip elevation(s) based on geotechnical design
requirements including the effects of scour, downdrag, or liquefaction

Obtain and verify final pile tip elevations and required resistances (factored and
unfactored loads) from the structural designer, finalize required pile tip elevations and
assess the following:

o
Determine the need to perform a pile drivability analysis to obtain required
tip elevation
o
Evaluate pile group settlement (if applicable). If settlement exceeds
allowable criteria, adjust pile lengths or the size of the pile layout and/or lengths

8.12.1

Determine the need for pile tip reinforcement

Required Pile Tip Elevation

Required pile tip elevations should typically be provided for all pile foundation design projects. The
required pile tip elevation is provided to ensure the constructed foundation meets the design
requirements of the project, which may include any or all of the following conditions and criteria:

Pile tip reaches the designated bearing layer

Scour

Downdrag

Uplift

Lateral loads

A general note is included on the bridge plans designating the Pile Tip Elevation for Minimum
Penetration for each bent.
The required tip elevation may require driving into, or through, very dense soil layers resulting in
potentially high driving stresses. Under these conditions a wave equation driveability analysis is
necessary to make sure the piles can be driven to the required embedment depth (tip elevation).
Higher grade steel (ASTM A252, Grade 3 or A572, Grade 50) are sometimes specified if needed to
meet driveability criteria. If during the structural design process, adjustments in the required tip

Volume 1
14

ODOT Geotechnical Design Manual


April 2010

elevations are necessary, or if changes in the pile diameter are necessary, the geotechnical designer
should be informed so that pile drivability can be re-evaluated.

8.12.2

Pile Drivability Analysis and Wave Equation Usage

High pile stresses often occur during pile driving operations and, depending on subsurface and
loading conditions, a Wave Equation analysis should always be considered to evaluate driving
stresses and the probability of pile damage. A pile driveability analysis is typically used in most pile
foundation designs to determine the nominal geotechnical resistance that a pile can be driven to
without damage. Foundation piles should typically be driven to the highest geotechnical axial
resistance feasible based on wave equation analysis so the maximum structural resistance of the pile
is utilized, resulting in the most cost-effective pile design.
All piles driven to nominal resistances greater than 600 kips should be driven based on wave
equation criteria. Piles driven to nominal resistance less than or equal to 600 kips may also require a
wave equation analysis depending on the subsurface conditions (such as very short end bearing
piles) and the pile loads. Engineering judgment is required in this determination. Pile driving stresses
should be limited to those described in AASHTO Article 10.7.8.

8.12.3

Pile Setup and Restrike

Using a waiting period and restrike after initial pile driving may be advantageous in certain soil
conditions to optimize pile foundation design. After initially driving the piles to a specified tip elevation,
the piles are allowed to set up for a specified waiting period, which allows pore water pressures to
dissipate and soil strength to increase. The piles are then restruck to confirm the required nominal
resistance.
The length of the waiting period depends primariiy on the strength and drainage characteristics of the
subsurface soils (how quickly the soil can drain) and the required nominal resistance. The minimum
waiting period specified in the Standard Specifications is 24 hours. If needed, this waiting period may
be extended in the contract special provisions to provide additional time for the soils to gain strength
and the piles to gain resistance. However, consideration should be given to increased contractor
standby costs that may be incurred by extended waiting periods. The pile design should compare the
cost and risk of extending the standard waiting period to gain sufficient strength versus designing and
driving the piles deeper to achieve the required bearing.
For projects with piles that require restrike, at least 1 pile per bent or 1 in 10 piles in a group
(whichever is more) should typically be restruck for pile acceptance. Additional restrike verification
testing should be conducted on any piles that indicate lower resistance at the end of initial driving or if
subsurface conditions vary substantially within a pile group. Restrike should be performed using a
warm pile hammer.
If pile acceptance from restrike is based on measured blow counts (dynamic formula or wave
equation methods), the restrike resistance (blow count) should be determined by measuring the total
pile set in the first 5 blows of driving and in successive 5 blow increments thereafter up to a total of at
least 20 blows or until refusal driving conditions are reached (>20 blows per inch). The driving
resistance reported (in blows per inch) is then determined by taking the inverse of the set per each 5
blow increment.

8.12.4

Driven Pile Types, and Sizes

The pile types generally used on most permanent structures are steel pipe piles (driven both open
and closed-end) and steel H-piles. Either H-pile or open-end steel pipe pile can be used for end
bearing conditions. For friction piles, steel pipe piles are often preferred because they can be driven
Volume 1
15

ODOT Geotechnical Design Manual


April 2010

closed-end (as full displacement piles) and because of their uniform cross section properties, which
provides the same structural bending capacity in any direction of loading. This is especially helpful
under seismic loading conditions where the actual direction of lateral loading is not precisely known.
Uniform section properties also aid in pile driving. Pipe piles are available in a variety of diameters
and wall thickness; however there are some sizes that are much more common than others and
therefore usually less expensive. The most common pipe pile sizes used on ODOT projects are:

PP 12.75 x 0.375

PP 16 x 0.500

PP 20 x 0.500

PP 24 x 0.500

Timber piles are occasionally used for temporary detour structures and occasionally on specialty
bridges, for retrofit or repair, and, on rare occasions, "in-kind" widening projects. ODOT standard
prestressed concrete piles are rarely used due to the following reasons:

They typically have less bending capacity than steel piles for a given size

They are difficult to connect to the pile cap for uplift resistance

They are inadequately reinforced for plastic hinge formation

Pile driving damage potential

Splicing difficulties

Cost, (typically more expensive than steel for a given capacity)

Prestressed concrete piles may however be appropriate in some areas like low seismic zones or
highly corrosive environments. The use of prestressed concrete piles is not prohibited in ODOT if
they are properly designed and cost effective.

The typical ASTM steel specifications and grades used in ODOT are as follows:

Steel Pipe Piles: ASTM A 252, Grade 2 & 3

Steel H-piles: ASTM A 572, Grade 50

The higher-grade steel may be required in some cases due to predicted high driving stresses or due
to high lateral bending stresses. Refer to AASHTO for ASTM requirements for other pile types.
Reinforced pile tips may be warranted in some cases where piles may encounter, or are required to
penetrate through, very dense cobbles and/or boulders. Pile tips are useful in protecting the tip of the
pile from damage. However, installing a reinforced pile tip does not eliminate all potential for pile
damage. High driving stresses may occur at these locations and still result in pile damage located
just above the reinforce pile tip. A driveability analysis should be performed in these cases where high
tip resistance is anticipated. All reinforced tips are manufactured from high strength (A27) steel.
Tip reinforcement for H-piles are typically called pile points. These come in a variety of shapes and
designs. H-pile tips are listed on the ODOT QPL. For pipe piles tip reinforcement are typically termed
shoes, although close-end points, like conical points, are also available. Pipe pile shoes may be
either inside or outside-fit. Besides protecting the pile tip, inside-fit shoes are sometimes specified to
help in delaying the formation of a pile plug inside the pipe pile so the pile may penetrate further
into, or even through, a relatively thin dense soil layer. If outside-fit shoes are specified, the outside lip
of the shoe may affect (reduce) the pile skin friction and this effect should be taken into account in the
pile design.

Volume 1
16

ODOT Geotechnical Design Manual


April 2010

8.12.5

Extreme Event Limit State Design

For the applicable factored loads for each extreme event limit state, the pile foundations shall be
designed to have adequate factored axial and lateral resistance.

8.12.5.1

Scour Effects on Pile Design

The effects of scour, where scour can occur, shall be evaluated in determining the required pile
penetration depth. The pile foundation shall be designed so that the pile penetration after the design
scour events satisfies the required nominal axial and lateral resistance. The pile foundation shall also
be designed to resist debris loads occurring during the flood event in addition to the loads applied
from the structure. The resistance factors for scour shall be those described in Section 8.10. The
axial resistance of the material lost due to scour should be determined using a static analysis. The
axial resistance of the material lost due to scour should not be factored. The piles will need to be
driven to the required nominal axial resistance plus the skin friction resistance that will be lost due to
scour.
From Equation 8-1:
i Qi Rn

(8-1)

The summation of the factored loads (iQi) must be less than or equal to the factored resistance
(Rn). Therefore, the nominal resistance needed, Rn, must be greater than or equal to the sum of the
factored loads divided by the resistance factor :
Rn (iQi)/dyn
For scour conditions, the resistance that the piles must be driven to needs to account for the
resistance in the scour zone that will not be available to contribute to the resistance required under
the extreme event (scour) limit state. The total driving resistance, Rndr, needed to obtain Rn, is
therefore:
Rndr = Rn + Rscour
Note that Rscour remains unfactored in this analysis to determine Rndr.
Pile design for scour is illustrated further in Figure 8 -15, where,

Rscour

= skin friction which must be overcome during driving through scour zone

Qp

= (iQi) = factored load per pile (KIPS)

(KIPS)

Dest.
pile (FT)

dyn

= estimated pile length needed to obtain desired nominal resistance per


= resistance factor

Volume 1
17

ODOT Geotechnical Design Manual


April 2010

Figure 8-15. Design of pile foundations for scour

Volume 1
18

ODOT Geotechnical Design Manual


April 2010

Seismic Design for Pile Foundations


For seismic design, all soil within and above liquefiable zones, shall not be considered to contribute
axial compressive resistance. Downdrag resulting from liquefaction induced settlement shall be
determined as specified in AASHTO and included in the loads applied to the foundation. Static
downdrag loads should not be combined with seismic downdrag loads due to liquefaction.
In general, the available factored geotechnical resistance should be greater than the factored loads
applied to the pile, including the downdrag, at the extreme event limit state. The pile foundation shall
be designed to structurally resist the downdrag plus structure loads. Pile design for liquefaction
downdrag is illustrated in Figure 8 -16, where,
RSdd
=
downdrag zone

skin friction which must be overcome during driving through

Qp

(iQi) = factored load per pile, excluding downdrag load

DD

downdrag load per pile

Dest.
per pile

estimated pile length needed to obtain desired nominal resistance

seis

resistance factor for seismic conditions

load factor for downdrag

The nominal bearing resistance of the pile needed to resist the factored loads, including downdrag, is
therefore,

Rn = (iQi)/seis + pDD/seis

The total driving resistance, Rndr, needed to obtain Rn, must account for the skin friction that has to be
overcome during pile driving that does not contribute to the design resistance of the pile. Therefore:

Rndr = Rn + RSdd

Note that RSdd remains unfactored in this analysis to determine Rndr.

Volume 1
19

ODOT Geotechnical Design Manual


April 2010

Figure 8-16. Design of pile foundations for liquefaction downdrag (WSDOT, 2006)
In the instance where it is not possible to obtain adequate geotechnical resistance below the lowest
layer contributing to downdrag (e.g., friction piles) to fully resist the downdrag, or if it is anticipated
that significant deformation will be required to mobilize the geotechnical resistance needed to resist
the factored loads including the downdrag load, the structure should be designed to tolerate the
settlement resulting from the downdrag and the other applied loads.
The static analysis procedures in AASHTO should be used to estimate the skin friction within, above
and below, the downdrag zone and to estimate pile lengths required to achieve the required bearing
resistance. For this calculation, it should be assumed that the soil subject to downdrag still
contributes overburden stress to the soil below the downdrag zone.
The pile foundation shall also be designed to resist the horizontal force resulting from lateral
spreading, if applicable, or the liquefiable soil shall be improved to prevent liquefaction and lateral
spreading. For lateral soil resistance of the pile foundation, the P-y curve soil parameters should be
reduced to account for liquefaction. To determine the amount of reduction, the duration of strong
shaking and the ability of the soil to fully develop a liquefied condition during the period of strong
shaking should be considered.
The force resulting from lateral spreading should be calculated as described in Chapter 6. In
general, the lateral spreading force should not be combined with the seismic forces. See Chapter 6 ,
Seismic Design for additional guidance regarding this issue.

Volume 1
20

ODOT Geotechnical Design Manual


April 2010

Drilled Shaft Foundation Design

Refer to AASHTO Article 10.8 for drilled shaft design requirements. Common shaft sizes range from
3 feet to 8 feet in diameter in 6 inch increments. Larger shaft diameters are also possible. The
minimum shaft diameter is 12 inches.
Once the shaft locations and foundation materials and properties are known, the drilled shaft design
process for normal bridge projects typically consists of the following:

Determine scour depths (if applicable),

Determine liquefaction potential and depths; estimate seismic induced settlement


(if applicable),

Evaluate long-term embankment settlement and downdrag potential,

Select most appropriate shaft diameter(s) in consultation with structure designer,

Determine (in consult with the structure designer) whether or not permanent
casing will be used,

Calculate nominal single shaft resistance as a function of depth,

Select and apply resistance factors to nominal resistance

Estimate downdrag loads (if applicable),

Estimate shaft or shaft group settlement and adjust shaft diameter or lengths if
necessary to limit settlement to service state limits,

Determine p-y curve parameters for lateral load analysis; modify parameters for
liquefied soils (if applicable),

The diameter of shafts will usually be controlled by the superstructure design loads and the
configuration of the structure but consideration should also be given to the foundation materials to be
excavated. If boulders or large cobbles are anticipated, attempt to size the shafts large enough so
the boulders or cobbles can be more easily removed if possible. Shaft diameters may also need to
be increased to withstand seismic loading conditions. The geotechnical engineer and the bridge
designer should confer and decide early on in the design process the most appropriate shaft
diameter(s) to use for the bridge, given the loading conditions, subsurface conditions at the site and
other factors. Also decide early on with the bridge designer if permanent casing is desired since this
will affect both structural and geotechnical designs. Specify each shaft as either a friction or end
bearing shaft since this dictates the final cleanout requirements in the specifications.
When the drilled shaft design calls for a specified length of shaft embedment into a bearing layer
(rock socket) and the top of the bearing layer is not well defined, an additional length of shaft
reinforcement should be added to the length required to reach the estimated tip elevation. This extra
length is required to account for the uncertainty and variability in the final shaft length. This practice is
much preferred instead of having to splice on additional reinforcement in the field during which time
the shaft excavation remains open. Any extra reinforcement length that is not needed can be easily
cut off prior to steel placement once the final shaft tip elevation is known. CSL tubes would also need
to be either cut off and recapped or otherwise adjusted. This additional reinforcement length should
be determined by the geotechnical engineer based on an evaluation of the site geology, location of
borehole information and the potential variability of the bearing layer surface at the plan location off
the shaft. The additional recommended length should be provided in the Geotechnical Report and
included in the project Special Provisions. Refer to the Standard Special Provisions for Section
00512 for further guidance and details of this application.
Volume 1
21

ODOT Geotechnical Design Manual


April 2010

If a minimum rock embedment (socket) depth is required, specify the reason for the rock
embedment. Try to minimize hard rock embedment depths if possible since this adds substantially to
the cost of drilled shafts.
Settlement may control the design of drilled shafts in cases where side resistance (friction) is
minimal, loads are high and the shafts are primarily end bearing on compressible soil. The shaft
settlement necessary to mobilize end bearing resistance may exceed that allowed by the bridge
designer. Confer with the bridge designer to determine shaft service loads and allowable amounts of
shaft settlement. Refer to the AASHTO methods to calculate the settlement of individual shafts or
shaft groups. Compare this settlement to the maximum allowable settlement and modify the shaft
design if necessary to reduce the estimated settlement to acceptable levels.

8.13.1

Nearby Structures

Where shaft foundations are placed adjacent to existing structures, the influence of the existing
structure(s) on the behavior of the foundation, and the effect of the foundation on the existing
structures, including vibration effects due to casing installation, should be investigated. In addition,
the impact of caving soils during shaft excavation on the stability of foundations supporting adjacent
structures should be evaluated. At locations where existing structure foundations are adjacent to the
proposed shaft foundation, or where a shaft excavation cave-in could adversely affect an existing
foundation, the design should require that casing be advanced as the shaft excavation proceeds.

8.13.2

Scour

The effect of scour shall be considered in the determination of the shaft penetration. The shaft
foundation shall be designed so that the shaft penetration and resistance remaining after the design
scour events satisfies the required nominal axial and lateral resistance. For this calculation, it shall be
assumed that the soil lost due to scour does not contribute to the overburden stress in the soil below
the scour zone. The shaft foundation shall be designed to resist debris loads occurring during the
flood event in addition to the loads applied from the structure.
Resistance factors for use with scour are described in Section 8.9.2. The axial resistance of the
material lost due to scour shall not be included in the shaft resistance.

8.13.3

Extreme Event Limit State Design of Drilled Shafts

The provisions of Section 8.12.5 shall apply. The nominal shaft resistance available to support
structure loads plus downdrag shall be estimated by considering only the positive skin and tip
resistance below the lowest layer contributing to the downdrag. For this calculation, it shall be
assumed that the soil contributing to downdrag does contribute to the overburden stress in the soil
below the downdrag zone.
In general, the available factored geotechnical resistance should be greater than the factored loads
applied to the shaft, including the downdrag, at the extreme limit state. The shaft foundation shall be
designed to structurally resist the downdrag plus structure loads.
In the instance where it is not possible to obtain adequate geotechnical resistance below the lowest
layer contributing to downdrag (e.g., friction shafts) to fully resist the downdrag, the structure should
be designed to tolerate the settlement resulting from the downdrag and the other applied loads.

Volume 1
22

ODOT Geotechnical Design Manual


April 2010

Micropiles

Micropiles shall be designed in accordance with Article 10.9 of the AASHTO LRFD Bridge Design
Specifications. Additional information on micropile design may be found in the FHWA Reference
Manual; Micropile Design and Construction (Publication No. FHWA NHI-05-039).

References

American Association of State Highway and Transportation Officials (AASHTO), 2010. AASHTO
LRFD Bridge Design Specifications, Customary U.S. Units. 5th Edition, with 2010 Interim Revisions.
AASHTO, 1988, Manual on Subsurface Investigations.
Bowles, J., 1988. Foundation Analysis and Design, 4th Edition, McGraw-Hill Book Company.
Briaud, J., and Tucker, L., 1997. Design and Construction Guidelines for Downdrag on Uncoated
and Bitumen-Coated Piles, NCHRP Report 393, Transportation Research Board, National Research
Council, Washington, D.C.
Briaud, J., 1989. The Pressuremeter for Highway Applications, FHWA-IP-89-008, Federal Highway
Administration, Washington, D.C.
Briaud, J. and Miran, J., 1992. The Cone Penetrometer Test, FHWA-IP-91-043, Federal Highway
Administration, Washington, D.C.
Samtani, N. and Nowatzki, E. 2006. Soils and Foundations Reference Manual, Volumes I and II,
FHWA-NHI-06-088, Washington, DC.
Elias, V., Christopher, B.R., and Berg, R.R., 2001, Mechanically Stabilized Earth Walls and
Reinforced Soil Slopes, Design and Construction Guidelines, Federal Highway Administration,
FHWA-NHI-00-043 (FHWA, 2001).
Gifford, D. G., J. R. Kraemer, J. R. Wheeler, and A. F. McKown. 1987. Spread Footings for Highway
Bridges. FHWA/RD-86/185. Federal Highway Administration, U.S. Department of Transportation,
Washington, DC, p. 229.
Hannigan, P.J., G.G. Goble, G. Thendean, G.E. Likins and F. Rausche, 1997. Design and
Construction of Driven Pile Foundations, Vol 1 & Vol II ,. FHWA-HI-97-03, Federal Highway
Administration, Washington, D.C., 822 pp.
Kavazanjian, E., Jr., Matasovi, T. Hadj-Hamou and Sabatini, P.J. 1997. Geotechnical Engineering
Circular No. 3, Design Guidance: Geotechnical Earthquake Engineering for Highways, Report No.
FHWA-SA-97-076, Federal Highway Administration, Washington, D.C.
Kimmerling, R. E. 2002. Geotechnical Engineering Circular No. 6, Shallow Foundations,
Report No. FHWA-SA-02-054, Federal Highway Administration, Washington, D.C.
Moulton, L. K., H. V. S. GangaRao, and G. T. Halverson. 1985. Tolerable Movement Criteria for
Highway Bridges, FHWA/RD-85/107. Federal Highway Administration, U.S. Department of
Transportation, Washington, DC, p. 118.

Volume 1
23

ODOT Geotechnical Design Manual


April 2010

ONeill, M. W. and Reese, L. C. 1999, Drilled Shafts: Construction Procedures and Design
Methods, Report No. FHWA-IF-99-025, Federal Highway Administration, Washington, D.C.
Oregon Department of Transportation, Bridge Design and Drafting Manual, Bridge Engineering
Section, 2004.
Oregon Department of Transportation, Hydraulics Manual, Geo-Environmental Section, 2005.
Oregon Department of Transportation, ODOT Soil and Rock Classification Manual ", GeoEnvironmental Section, 1987.
Oregon Department of Transportation, "Standard Specifications for Highway Construction", 2002
Edition and related Standard Special Provisions.
Reese, L. C. 1984. Handbook on Design of Piles and Drilled Shafts Under Lateral Load. FHWA-IP84/11, Federal Highway Administration, U.S. Department of Transportation, Washington, DC.
Reese, L. C., 1986. Behavior of Piles and Pile Groups Under Lateral Load,,Report No. FHWA/RD85/106, U. S. Department of Transportation, Federal Highway Administration, Office of Engineering
and Highway Operations Research and Development, Washington D. C., 311
Richardson, E. V. and Davis, S. R. 2001. Hydraulic Engineering Circular No. 18, Evaluating Scour
at Bridges, 4th Edition, FHWA-NHI-01-001 HEC-18, Federal Highway Administration, Washington,
D.C.
Sabatini, P.J., Bachus, R.C., Mayne, P.W., Schneider, J.A., and Zettler, T.E., 2002, Geotechnical
Engineering Circular No. 5, Geotechnical Engineering Circular No. 5, Evaluation of Soil and
Rock Properties, U.S. Department of Transportation, Federal Highway Administration, Washington,
D.C.
Sabatini, P.J., Tanyu, B., Armour, T., Groneck, P., Keeley, J., 2005. Micropile Design and
Construction (Reference Manual for NHI Course 132078), Federal Highway Administration Report
No. FHWA NHI-05-039, Washington, D.C.
Samtani, N.C., Nowatzki, E.A., Mertz, D.R., 2010. Selection of Spread Footings on Soil to Support
Highway Bridge Structures, Federal Highway Administration Report No. FHWA-RC/TD-10-001,
Washington, D.C.
Urzua, A., 1996. CBEAR: Bearing Capacity of Shallow Foundations Users Manual, FHWA-SA-94034, Federal Highway Administration, Washington, D.C.
Urzua, A., 1993. EMBANK: A Microcomputer Program to Determine One-Dimensional Compression
Settlement Due to Embankment Loads, Users Manual, FHWA-SA-92-045, Federal Highway
Administration, Washington, D.C.

Volume 1
24

ODOT Geotechnical Design Manual


April 2010

Chapter

Embankments Analysis and


Design

General

This chapter addresses the analysis and design of rock and earth embankments. Also addressed
briefly are foundation improvement (ground improvement), the use of lightweight fill and settlement
and stability mitigation techniques. Bridge approach embankments, defined as fill under bridge ends,
are not covered in this chapter, but are addressed in Chapter 8 and in Chapter 6. The primary
geotechnical issues that impact embankment performance are overall (global) stability, internal
(slope) stability, settlement, material selection, compaction, and constructability. For the purposes of
this chapter, embankments include the following:

Rock embankments, also known as all-weather embankments, are defined as fills in which
the material is non-moisture-density testable and is composed of durable granular materials.

Earth embankments are fills that are typically composed of onsite or imported borrow, and
could include a wide variety of materials from fine to coarse grain. The material is usually
moisture-density testable.

Lightweight fills contain lightweight fill or recycled materials as a significant portion of the
embankment volume, and the embankment construction is specified by special provision.
Lightweight fills are most often used as a portion of the bridge approach embankment to
mitigate settlement or in landslide repairs to reestablish roadways.

Embankments under 10 feet high in areas of stable ground and with slopes flatter than 2:1 generally
do not require a detailed geotechnical investigation and analysis. These embankments can be
designed based on past experience with similar soils and on engineering judgment. Embankments
over 10 feet high, with steeper slopes, constructed in problem soil areas, or from specially designed
or unique materials will require a detailed geotechnical analysis, development of special provisions
and possibly details included in the contract plans.

Volume 1
1

ODOT Geotechnical Design Manual


April 2010

Design Considerations

9.2.1

Typical Embankment Materials and Compaction

New embankments and embankment widenings require suitable fill materials be used and properly
compacted with the right equipment for the type of material. Compaction control of soil
embankments requires the development of moisture-density relationships to allow measurement of
in-place compaction during construction. Tamping foot rollers and specified passes of the rollers are
used to achieve the required density of the fill. Non-durable rock materials may require additional
compactive effort beyond the usual soil construction methods to prevent long term settlement of an
embankment. The ODOT Standard Specifications for Construction identifies the acceptable
embankment construction methods for soil, non-durable rock and rock materials. The geotechnical
designer should determine during the exploration program if any of the material from planned
earthwork excavations will be suitable for embankment construction. Consideration should be given
as to whether the material is moisture sensitive and difficult to compact during wet weather.

9.2.1.1

Al l - Wea t h e r E m b an k m e n t M a t e r i a l s

ODOT projects are increasingly being constructed within shorter time frames that may require
fill placement occurring at any time of the year. Clean, granular, all-weather embankment
materials allow the contractor the ability to properly place and compact fill materials year round.
Clean, granular fill material use also provides better access to work areas, and facilitates
construction staging and traffic detouring. The ODOT Standard Specifications identify 2
materials considered to be suitable for all-weather construction: Selected Stone Backfill (section
00330.15), and Stone Embankment Material (section 00330.16). Both of these materials have
in common the use of durable material, as defined in section 00110.20 Durable Rock.
Compaction tests cannot be applied to coarse material with any degree of accuracy; therefore, a
method specification approach is typically specified for granular embankments, as described in
section 00330.43 Non-Moisture Density Testable Materials.

9.2.1.2

D u r a b l e a n d N o n - Du r a b l e Ro c k M at e r i a l s

Special consideration should be given during design to the type of material that will be used in rock
embankments. In some areas of the state, moderately weathered or very soft rock may be
encountered in cuts and used as embankment fill. Follow these guidelines:

Degradable fine grained sandstone and siltstone are often encountered in the cuts and the
use of this material in embankments can result in significant long term settlement and stability
problems as the rock degrades, unless properly compacted with heavy tamping foot rollers
(Machan, et al., 1989). The slake durability test (ASTM D4644) should be performed if the
geologic nature of the rock source proposed indicates that poor durability rock is likely to be
encountered.

When the rock is found to be non-durable, it should be physically broken down and
compacted as earth embankment provided the material meets or exceeds common borrow
requirements. Special compaction requirements, defined by method specification, may be
needed for these materials. In general, tamping foot rollers work best for breaking down the
rock fragments. The minimum size roller should be about 30 tons. Specifications should
include the maximum size of the rock fragments and maximum lift thickness. These
requirements will depend on the hardness of the rock, and a test section should be

Volume 1
2

ODOT Geotechnical Design Manual


April 2010

incorporated into the contract to verify that the Contractors methods will achieve compaction
and successfully break down the material. In general, both the particle size and lift thickness
should be limited to 12 inches.

9.2.1.3

E a r t h E m b a n k m en t s

Embankments constructed with common borrow materials must be placed in accordance with the
procedures of the ODOT Standard Specifications, section 00330 Earthwork. These specifications
are intended for use where it is not necessary to strictly control the strength properties of the
embankment material and where all-weather construction is not required.

9.2.2

Embankment Stability Assessment

In general, embankments 10 feet or less in height with 2H:1V or flatter side slopes, may be designed
based on past precedence and engineering judgment provided there are no known problem soil
conditions such as liquefiable sands, organic soils, soft/loose soils, or potentially unstable soils such
as clay, estuarine deposits, or peat. Embankments over 10 feet in height or any embankment on soft
and/or unstable soils or those comprised of light weight fill require more in depth stability analyses, as
do any embankments with side slope inclinations steeper than 2H:1V. Moreover, any fill placed near
or against a bridge abutment or foundation, or that can impact a nearby buried or above-ground
structure, will likewise require stability analyses by the geotechnical designer. Slope stability analysis,
discussed in Chapter 7, are to be conducted in accordance with the standard of practice for
geotechnical engineering.
The geotechnical designer should determine key issues that need to be addressed to perform
stability analysis. These include:

Is the site underlain by soft silt, clay or peat? If so, a staged stability analysis (staged
construction of fill with stability analysis at each stage) may be required.

Are site constraints such that slopes steeper than 2H:1V are required? If so, a detailed slope
stability assessment is needed to evaluate the various alternatives.

Is the embankment temporary or permanent? Factors of safety for temporary embankments


may be lower than for permanent ones, depending on the site conditions and the potential for
variability.

Will the new embankment impact nearby structures or bridge abutments? If so, more
thorough sampling, testing and analysis are required.

Are there potentially liquefiable soils at the site? If soil, seismic analysis to evaluate this may
be warranted see Chapter 6 and ground improvement may be needed.

Several methodologies for analyzing the stability of slopes are detailed or identified by reference in
Chapter 7 and are directly applicable to earth embankments.

9.2.2.1

Safety Factors

All embankments not supporting or potentially impacting structures shall have a minimum safety
factor of 1.25. Embankments supporting or potentially impacting non-critical structures shall have a
minimum factor of safety of 1.3. As discussed in Section 8.7, all bridge approach embankments and
embankments supporting critical structures shall have a safety factor of 1.5.

Volume 1
3

ODOT Geotechnical Design Manual


April 2010

Under seismic conditions, only those portions of the new embankment that could impact an adjacent
structure such as bridge abutments and foundations or nearby buildings require seismic analyses
and an adequate overall stability resistance factor (i.e., a maximum resistance factor of 0.9 or a
minimum factor of safety of 1.1). See Chapter 6 for specific requirements regarding seismic design
of embankments.

9.2.2.2

Strength Parameters

Strength parameters are required for any stability analysis. Strength parameters appropriate for the
different types of stability analyses are determined based on Chapter 5 and by reference to FHWA
Geotechnical Engineering Circular No. 5 (Sabatini, et al., 2002). If the critical stability is under drained
conditions, such as in sand or gravel, then effective stress analysis using a peak friction angle is
appropriate and should be used for stability assessment. In the case of over-consolidated fine
grained soils, a friction angle based on residual strength may be appropriate. This is especially true
for soils that exhibit strain softening or are particularly sensitive to shear strain. If the critical stability is
under undrained conditions, such as in most clays and silts, a total stress analysis using the
undrained cohesion value with no friction is appropriate and should be used for stability assessment.
For staged construction, both short (undrained) and long term (drained) stability need to be
assessed. At the start of a stage the input strength parameter is the undrained cohesion. The total
shear strength of the fine-grained soil increases with time as the excessive pore water dissipates,
and friction starts to contribute to the strength.

9.2.3

Embankment Settlement Assessment

New embankments and embankment widenings should be analyzed using the methods discussed in
the FHWA Soils and Foundation Reference Manual, (Samtani, N. and Nowatzki, E. 2006).
Laboratory test results of foundation soil samples obtained should be used as a basis for determining
the primary and secondary settlement amounts and rates. Because primary consolidation and
secondary compression can continue to occur long after the embankment is constructed (post
construction settlement), they represent the major settlement concerns for embankment design and
construction. Post construction settlement can damage structures and utilities located within the
embankment, especially if those facilities are also supported by adjacent soils or foundations that do
not settle appreciably, leading to differential settlements. If the primary consolidation is allowed to
occur prior to placing utilities or building structures that would otherwise be impacted by the
settlement, the impact is essentially mitigated. However, it can take weeks to years for primary
settlement to be essentially complete, and significant secondary compression of organic soils can
continue for decades. Many construction projects cannot absorb the scheduling impacts associated
with waiting for primary consolidation and/or secondary compression to occur. Therefore, estimating
the time rate of settlement is often as important as estimating the magnitude of settlement.

9.2.3.1

S e t t l e m e n t An a l ys i s

The key parameters for evaluating the amount of settlement below an embankment include
knowledge of:

The subsurface profile including soil types, layering, groundwater level and unit weights;

The compression indexes for primary, rebound and secondary compression from laboratory
test data, correlations from index properties, and results from settlement monitoring programs
completed for the site or nearby sites with similar soil conditions.

Volume 1
4

ODOT Geotechnical Design Manual


April 2010

The geometry of the proposed fill embankment, including the unit weight of fill materials and
any long term surcharge loads.

9.2.3.2

An a l y t i c a l Too l s

The primary consolidation and secondary settlement can be calculated by hand or by using computer
programs such as EMBANK (FHWA, 1993). Alternatively, spreadsheet solutions can be easily
developed. The advantage of computer programs such EMBANK are that multiple runs can be made
quickly, and they include subroutines to estimate the increased vertical effective stress caused by the
embankment or other loading conditions.

Stability Mitigation

A variety of techniques are available to mitigate inadequate slope stability for new embankments
or embankment widenings. These techniques include staged construction to allow for the
underlying soils to gain strength, base reinforcement, ground improvement, use of lightweight
fill, and construction of toe berms (counterweights) and shear keys. A summary of these
instability mitigation techniques is presented below along with the key design considerations.

9.3.1

Staged Construction

Where soft compressible soils are present below a new embankment location and it is not
economical to remove and replace these soils with compacted fill, the embankment can be
constructed in stages to allow the strength of the compressible soils to increase under the weight of
new fill. Construction of the second and subsequent stages commences when the strength of the
compressible soils is sufficient to maintain stability. In order to define the allowable height of fill for
each stage and maximum rate of construction, detailed geotechnical analysis is required. The
analysis to define the height of fill placed during each stage and the rate at which the fill is placed is
typically completed using a limit equilibrium slope stability program along with time rate of settlement
analysis to estimate the percent consolidation required for stability. Field monitoring of settlement
and pore water pressures are usually required during construction.

9.3.2

Base Reinforcement

Base reinforcement may be used to increase the factor of safety against slope failure. Base
reinforcement typically consists of placing a geotextile or geogrid at the base of an embankment prior
to constructing the embankment. Base reinforcement is particularly effective where soft/weak soils
are present below a planned embankment location. The base reinforcement can be designed for
either temporary or permanent applications. Most base reinforcement applications are temporary, in
that the reinforcement is needed only until the underlying soils shear strength has increased
sufficiently as a result of consolidation under the weight of the embankment, see Section 9.3.1.
Therefore, the base reinforcement does not need to meet the same design requirements as
permanent base reinforcement regarding creep and durability. The design of base reinforcement is
similar to the design of a reinforced slope in that limit equilibrium slope stability methods are used to
determine the strength required to obtain the desired safety factor. The detailed design procedures
provided by Holtz, et al. (1995) should be used for embankments utilizing base reinforcement.
Base reinforcement materials should be placed in continuous longitudinal strips in the direction of
main reinforcement. Joints between pieces of geotextile or geogrid in the strength direction
(perpendicular to the slope) should be avoided. All seams in the geotextiles should be sewn and not
lapped. Likewise, geogrids should be linked with mechanical fasteners or pins and not simply
overlapped. Where base reinforcement is used, the use of Select Stone Backfill or Stone
Volume 1
5

ODOT Geotechnical Design Manual


April 2010

Embankment Material, instead of common or select borrow, may also be needed to increase the
embankment shear strength.

9.3.3

Ground Improvement

Refer to Chapter 11for references and information on ground improvement design.

9.3.4

Lightweight Fills

Lightweight embankment fill may be used to improve embankment stability. Lightweight fill materials
are generally used to reduce driving forces contributing to instability, and reduce potential settlement
resulting from consolidation of compressible foundation soils. Situations where lightweight fill may be
appropriate include conditions where the construction schedule does not allow the use of staged
construction, where existing utilities or adjacent structures are present that cannot tolerate the
magnitude of settlement induced by placement of typical fill, and at locations where post-construction
settlements may be excessive under conventional fills. Lightweight fill can consist of a variety of
materials including polystyrene blocks (geofoam), light weight aggregates (rhyolite, expanded shale,
blast furnace slag, fly ash), wood fiber, shredded rubber tires, and other materials. Lightweight fills
are infrequently used due to either high costs or other disadvantages with using these materials.

9.3.5

Toe Berms and Shear keys

Toe berms and shear keys are methods to improve the stability of an embankment by increasing the
resistance along potential failure surfaces. Toe berms are typically constructed of granular materials
that can be placed quickly, do not require much compaction, and have relatively high shear strength.
ODOT would typically specify the use of Stone Embankment Material when toe berms and shear
keys are required.

Settlement Mitigation

9.4.1

Acceleration Using Wick Drains

Wick drains, or prefabricated drains, are in essence, vertical drainage paths that can be installed into
compressible soils to decrease the overall time required for completion of primary consolidation. Wick
drain design considerations, example designs, guideline specifications, and installation
considerations are provided by reference in Chapter 11. Section 00435 of the ODOT Standard
Specifications addresses installation of wick drains.

9.4.2

Acceleration Using Surcharges

Surcharge loads are additional loads placed on the fill embankment above and beyond the finish
grades. The primary purpose of a surcharge is to speed up the consolidation process. Two
significant design and construction considerations for using surcharges include embankment stability
and re-use of the additional fill materials. New embankments over soft soils can result in stability
problems. Adding additional surcharge fill could exacerbate the stability problem. Furthermore, after
the settlement objectives have been met, the surcharge will need to be removed. If the surcharge
material cannot be moved to another part of the project site for use as site fill or as another
surcharge, it is often not economical to bring the extra surcharge fill to the site only to haul it away
again. Also, when fill soils must be handled multiple times (such as with a rolling surcharge), it is
advantageous to use gravel borrow to reduce workability issues during wet weather conditions.

Volume 1
6

ODOT Geotechnical Design Manual


April 2010

9.4.3

Lightweight Fills

Lightweight fills can also be used to mitigate settlement issues as indicated in Section 9.3.4.
Lightweight fills reduce the new loads imposed on the underlying compressible soils, thereby
reducing the magnitude of the settlement.

9.4.4

Subexcavation

Subexcavation refers to excavating the soft compressible or unsuitable soils from below the
embankment footprint and replacing these materials with higher quality, less compressible material.
Because of the high costs associated with excavating and disposing of unsuitable soils as well as the
difficulties associated with excavating below the water table, subexcavation and replacement typically
only makes economic sense under certain conditions. Some of these conditions include, but are not
limited to:

The area requiring overexcavation is limited;

The unsuitable soils are near the ground surface and do not extend very deep (typically, even
in the most favorable of construction conditions, subexcavation depths greater than about 10
ft are in general not economical);

Temporary shoring and dewatering are not required to support or facilitate the excavation
and;

Suitable materials are readily available to replace the over-excavated unsuitable soils.

Volume 1
7

ODOT Geotechnical Design Manual


April 2010

References

Federal Highway Administration, 1992, "EMBANK, Computer Program, User's Manual Publication ,
Publication No. FHWA-SA-92-045.
Holtz, R. D., Christopher, B. R., and Berg, R. R., 1995, Geosynthetic Design and Construction
Guidelines, Federal Highway Administration, FHWA HI-95-038.
Machan, G., Szymoniak, T. and Siel, B., 1989, Evaluation of Shale Embankment Construction
Criteria, Experimental Feature Final Report OR 83-02, Oregon State Highway Division,
GeotechnicalEngineering Group.
Sabatini, P.J, Bachus, R.C, Mayne, P.W., Schneider, J.A., Zettler, T.E. (2002), Geotechnical
Engineering Circular No. 5, Evaluation of Soil and Rock Properties, Report No FHWA-IF-02-034.
Samtani, N. and Nowatzki, E. (2006), Soils and Foundation Reference Manual, Volumes I and II,
Report No. FHWA NHI-06-088.

Volume 1
8

ODOT Geotechnical Design Manual


April 2010

Chapter

10

Soil Cuts - Analysis and Design

General

Soil cut slope design must consider many factors such as the materials and conditions present in the
slope, materials available or required for construction on a project, space available to make the
slopes, minimization of future maintenance and slope erosion. Soil slopes less 10 feet high are
generally designed based on past experience with similar soils and on engineering judgment. Cut
slopes greater than 6 to 10 feet in height usually require a more detailed geotechnical analysis.
Relatively flat (2H:1V or flatter) cuts in granular soil when groundwater is not present above the ditch
line, will probably not require rigorous analysis. Any cut slope where failure would result in large
rehabilitation costs or threaten public safety should obviously be designed using more rigorous
techniques. Situations that will warrant more in-depth analysis include:

large cuts,

cuts with irregular geometry,

cuts with varying stratigraphy (especially if weak zones are present),

cuts where high groundwater or seepage forces are likely,

cuts involving soils with questionable strength, or

cuts in old landslides or in formations known to be susceptible to landsliding.

A major cause of cut slope failure is related to reduced confining stress within the soil upon
excavation. Undermining the toe of the slope, increasing the slope angle, and cutting into heavily
overconsolidated clays have also resulted in slope failures. Careful consideration should be given to
preventing these situations by surcharging or buttressing the base of the slope, choosing an
appropriate slope angle (i.e., not oversteepening), and by keeping drainage ditches a reasonable
distance away from the toe of slope. Cutslopes in heavily overconsolidated clays may require special
mitigation measures, such as retaining walls rather than an open cut in order to prevent slope
deformation and reduction of soil strength to a residual value. Consideration should also be given to
establishing vegetation on the slope to prevent long-term erosion. It may be difficult to establish
vegetation on slopes with inclinations steeper than 2H:1V without the use of erosion mats or other
stabilization methods.
Volume 1
1

ODOT Geotechnical Design Manual


April 2010

10.1.1

Design Parameters

The major cutslope design parameters are slope geometry, soil shear strength and predicted or
measured groundwater levels. For cohesionless soil, stability of a cut slope is independent of height
and therefore slope angle becomes the key parameter of concern. For cohesive (= 0) soils, the
height of the cut becomes the critical design parameter. For c- and saturated soils, slope stability is
dependent on both slope angle and height of cut. Also critical to the proper design of cut slopes is the
incorporation of adequate surface and subsurface drainage facilities to reduce the potential for future
stability or erosional problems.
Establishment of design parameters is done by a thorough site reconnaissance, sufficient exploration
and sampling, and a laboratory testing program designed to identify the material soil strength
properties to be used in analysis. Backanalysis methods may also be used to determine the
appropriate shear strength for design. The geotechnical designer should be familiar with the state of
the practice in determining the design parameters for analysis. References are presented in
Section 10.3.

Soil Cut Design

10.2.1

Design Approach and Methodology

Safe design of cut slopes is typically based on past experience or on more in-depth analysis. Both
approaches require accurate sire specific information regarding geologic conditions obtained from
standard field and laboratory classification procedures. Design guidance for simple projects is
provided in the ODOT Highway Design Manual, located on the ODOT website, and can be used
unless indicated otherwise by the geotechnical designer. Slopes less than 6 to 10 feet high, with
slopes flatter than 2:1, may be used without in-depth analysis if no special concerns are noted by the
geotechnical designer. If the geotechnical designer determines that a slope stability study is
necessary, information that will be needed for analysis includes:

an accurate cross section showing topography,

proposed grade,

soil unit profiles,

unit weight and strength parameters (c,), (c,), or Su (depending on soil type and drainage
and loading conditions) for each soil unit, and

location of the water table and flow characteristics.

The design factor of safety for static slope stability is 1.25. This safety factor should be increased to a
minimum of 1.30 for slopes where failure would cause significant impact to adjacent structures. For
pseudo-static seismic analysis the factor of safety can be decreased to 1.1. Cut slopes are generally
not designed for seismic conditions unless slope failure could impact adjacent structures. These
factors of safety should be considered as minimum values. The geotechnical designer should decide
on a case by case basis whether or not higher factors of safety should be used based the
consequences of failure, past experience with similar soils, and uncertainties in analysis related to
site and laboratory investigation.

Volume 1
2

ODOT Geotechnical Design Manual


April 2010

Preliminary slope stability analysis can be performed using simple stability charts. See Abramson, et
al. (1994) for example charts. These charts can be used to determine if a proposed cut slope might
be subject to slope failure. If slope instability appears possible, or if complex conditions exist beyond
the scope of the charts, more rigorous computer methods such as XSTABL, PCSTABL, and
SLOPE/W can be employed see Chapter 7. Effective use of these programs requires accurate
determination of site geometry including surface profiles, soil unit boundaries, and location of the
water table, as well as unit weight and strength parameters for each soil type.

10.2.2

Seepage Analysis and Impact on Design

The introduction of groundwater to a slope is a common cause of slope failures. The addition of
groundwater often results in a reduction in the shear strength of soils. A higher groundwater table
results in higher pore pressures, causing a corresponding reduction in effective stress and soil shear
strength. A cutslope below the groundwater table results in destabilizing seepage forces, adds weight
to the soil mass, increasing driving forces for slope failures. It is important to identify and accurately
model seepage within proposed cut slopes so that adequate slope and drainage designs are
employed. suring the phreatic (water table) surface with open standpipes or observation wells.
Piezometric data from piezometers can be used to estimate the phreatic surface or piezometric
surface if confined flow conditions exist. A manually prepared flow net or a numerical method such as
finite element analysis can be used provided sufficient boundary information is available. The pore
pressure ratio (ru) can also be used. However, this method is generally limited to use with stability
charts or for determining the factor of safety for a single failure surface.

10.2.3

Surface and Subsurface Drainage Considerations


and Design

The importance of adequate drainage cannot be overstated when designing cut slopes. Surface
drainage can be accomplished through the use of drainage ditches and berms located above the top
of the cut, around the sides of the cut, and at the base of the cut. Surface drainage facilities should
direct surface water to suitable collection facilities.
Subsurface drainage should be employed to reduce driving forces and increase soil shear strength
by lowering the water table, thereby increasing the factor of safety against a slope failure. Subsurface
conditions along cut slopes are often heterogeneous. Thus, it is important to accurately determine the
geologic and hydrologic conditions at a site in order to place drainage systems where they will be the
most effective. Subsurface drainage techniques available include:

cut-off trenches (French drains)

horizontal drains

relief wells

Cut-off trenches: Cut-off trenches, also known as French drains, are a gravel filled trench near the
top of the cut slope to intercept groundwater and convey it around the slope. They are effective for
shallow groundwater depths from 2 to 15 feet deep.
Horizontal drains: If the groundwater table needs to be lowered to a greater depth, horizontal drains
can be installed, if the soils are noncohesive and granular in nature. Horizontal drains are generally
not very effective in finer grained soils. Horizontal drains consist of small diameter holes drilled at
slight angles into a slope face and backfilled with perforated pipe wrapped in drainage geotextile.
Installation might be difficult in soils containing boulders, cobbles or cavities. Horizontal drains require
periodic maintenance as they tend to become clogged over time.
Volume 1
3

ODOT Geotechnical Design Manual


April 2010

Relief wells: Relief wells can be used in situations where the water table is at a great depth. They
consist of vertical holes cased with perforated pipe connected to a disposal system such as
submersible pumps or discharge channels similar to horizontal drains. They are generally not
common in the construction of cut slopes.
Whatever subsurface drainage system is used, monitoring should be implemented to determine its
effectiveness. Typically, piezometers or observation wells are installed during exploration. These
should be left in place and periodic site readings should be taken to determine groundwater levels or
pore pressures depending on the type of installation. High readings would indicate potential problems
that should be mitigated before a failure occurs.
Surface drainage, such as brow ditches at the top of the slope, and controlling seepage areas as the
cut progresses and conveying that seepage to the ditch at the toe of the cut, should be applied to all
cut slopes. Subsurface drainage is more expensive and should be used when stability analysis
indicates pore pressures need to be lowered in order to provide a safe slope. The inclusion of
subsurface drainage for stability improvement should be considered in conjunction with other
techniques outlined below to develop the most cost effective design meeting the required factor of
safety.

10.2.4

Stability Improvement Techniques

There are a number of options that can be used in order to increase the stability of a cut slope.
Techniques include:

Flattening slopes

Benching slopes

Lowering the water table (discussed previously)

Structural systems such as retaining walls or reinforced slopes.

Changing the geometry of a cut slope is often the first technique considered when looking at
improving stability. For flattening a slope, enough right-of-way must be available. As mentioned
previously, stability in purely dry cohesionless soils depends on the slope angle, while the height of
the cut is often the most critical parameter for cohesive soils. Thus, flattening slopes usually proves
more effective for granular soils with a large frictional component.
Structural systems are generally more expensive than the other techniques, but might be the only
option when space is limited. Shallow failures and sloughing can be mitigated by placing a 2 to 3-foot
thick rock drainage blanket over the slope in seepage areas. Moderate to high survivability
permanent erosion control geotextile should be placed between native soil and drain rock to keep
fines from washing out and/or clogging the drain rock. In addition, soil bioengineering can be used to
stabilize cut slopes against shallow failures (generally less than 3 feet deep), surface sloughing and
erosion along cut faces.

Volume 1
4

ODOT Geotechnical Design Manual


April 2010

10.2.5

Erosion and Piping Considerations

Surface erosion and subsurface piping are most common in clean sands, nonplastic silts and
dispersive clays. Loess and volcanic ash are particularly susceptible. However, all cut slopes should
be designed with adequate drainage and temporary and permanent erosion control facilities to limit
erosion and piping as much as possible. The amount of erosion that occurs along a slope is a factor
of soil type, rainfall intensity, slope angle, length of slope, and vegetative cover. The first two factors
cannot be controlled by the designer, but the last three factors can. Longer slopes can be terraced at
approximate 15- to 30-foot intervals with drainage ditches installed to collect water. Best
Management Practices (BMPs) for temporary and permanent erosion and stormwater control as
outlined in the ODOT Highway Design Manual should always be used. Construction practices should
be specified that limit the extent and duration of exposed soil. For cut slopes, consideration should be
given to limiting earthwork during the wet season and requiring that slopes be covered as they are
exposed, particularly for the highly erodable soils mentioned above.

10.2.6

Sliver Cuts

A sliver cut is defined as slope excavation less than 10 feet wide over some or all of its height. Sliver
cuts in soils should be avoided because they are difficult to build. Cuts at least 10 feet wide over the
full height of the cut require the use of conventional earth moving machinery to maximize production.
Cuts less than 10 feet wide and up 25 feet high measured along the slope can be excavated with a
large backhoe but at the expense of production. If a sliver cut is used, consider how it will be built
and be sure to account for the difficulty in the cost estimate.

Volume 1
5

ODOT Geotechnical Design Manual


April 2010

References

Abramson, L., Boyce, G., Lee, T., and Sharma, S., 1994, Advanced Technology for Soil Slope
Stability,. Vol. 1, FHWA-SA-94-005.
FHWA, 1993, Soils and Foundations Workshop Manual, Second Edition, FHWA-HI-88-009.
ODOT Highway Design Manual, 2003.
Turner, A., and Schuster, R., 1996, Landslides: Investigation and Mitigation, TRB Special Report 247.

Volume 1
1

ODOT Geotechnical Design Manual


April 2010

Chapter

11

Ground Improvement

General

Ground improvement is used to address a wide range of geotechnical engineering problems,


including, but not limited to, the following:

Improvement of soft or loose soil to reduce settlement, increase bearing resistance, and/or to
improve overall stability of bridge foundations, retaining walls, and/or for embankments.

To mitigate liquefiable soils.

To improve slope stability for landslide mitigation.

To retain otherwise unstable soils.

To improve workability and usability of fill materials.

To accelerate settlement and soil shear strength gain.

Types of ground improvement techniques include the following:

Vibrocompaction techniques such as stone columns and vibroflotation, and other techniques
that use vibratory probes that may or may not include compaction of gravel in the hole
created to help densify the soil.

Deep dynamic compaction.

Blast densification.

Geosynthetic reinforcement of embankments.

Wick drains, sand columns, and similar methods that improve the drainage characteristics of
the subsoil and thereby help to remove excess pore water pressure that can develop under
load applied to the soil.

Grout injection techniques and replacement of soil with grout, such as compaction grouting,
jet grouting, and deep soil mixing.

Volume 1
1

ODOT Geotechnical Design Manual


April 2010

Lime or cement treatment of soils to improve their shear strength and workability
characteristics.

Permeation grouting and ground freezing (temporary applications only).

Each of these methods has limitations regarding their applicability and the degree of improvement
that is possible.

Development of Design Parameters


and Other Input Data for Ground
Improvement Analysis

In general, the geotechnical investigation conducted to design the cut, fill, structure foundation,
retaining wall, etc., that the improved ground is intended to support will be adequate for the design of
the soil improvement technique proposed. However, specific soil information may need to be
emphasized depending on the ground improvement technique selected. For example, for vibrocompaction techniques, deep dynamic compaction, and blast densification, detailed soil gradation
information is critical to the design of such methods, as minor changes in soil gradation
characteristics could affect method feasibility. Furthermore, the in-situ soil testing method used (e.g.,
SPT testing, cone testing, etc.) will need to correspond to the technique specified in the contract to
verify performance of the ground improvement technique, as the test data obtained during design will
be the baseline to which the improved ground will be compared. Other feasibility issues will need to
be addressed if these types of techniques are used. Ground vibrations caused by the improvement
technique may have critical impacts on adjacent structures. Investigation of the foundation and soil
conditions beneath adjacent structures and utilities may be needed, (in addition to standard
precondition surveys of the structures) to enable evaluation of the risk of damage caused by the
ground improvement technique.

Wick Drains: For wick drains, the ability to penetrate the soil with the wick drain mandrel, in
addition to obtaining rate-of-settlement information, must be assessed. Atterberg limit and
water content data should be obtained, as well as any other data that can be useful in
assessing the degree of overconsolidation of the soil present, if any.

Grout Injection Techniques: Grout injection techniques (not including permeation grouting)
can be used in a fairly wide range of soils, provided the equipment used to install the grout
can penetrate the soil. The key is to assess the ability of the equipment to penetrate the soil,
assign soil density and identify potential obstructions such as boulders.

Permeation Grouting: Permeation grouting is more limited in its application, and its
feasibility is strongly dependent on the ability of the grout to penetrate the soil matrix under
pressure. To evaluate the feasibility of these two techniques, detailed grain size
characterization and permeability assessment must be conducted, as well as the effect
groundwater may have on these techniques. An environmental assessment of such
techniques may also be needed, especially if there is potential to contaminate groundwater
supplies.

Ground Freezing: Similarly, ground freezing is a highly specialized technique that is strongly
depending on the soil characteristics and groundwater flow rates present.

Volume 1
2

ODOT Geotechnical Design Manual


April 2010

Design Requirements

The following design manuals and references shall be used for specific ground improvement
applications:

General Ground Improvement Design Requirements:


FHWA manual No. FHWA-SA-98-086, Ground Improvement Technical
Summaries, (Elias, et al., 2000)

Stone Column Design:


FHWA Report No. FHWA/RD-83/O2C, Design and Construction of Stone
Columns, (Barkdale and Bachus, 1983)

Deep Dynamic Compaction:


FHWA manual No. FHWA-SA-95-037, Geotechnical Engineering Circular No.
1, Dynamic Compaction, (Lukas, 1995)

Wick Drain Design:


FHWA manual FHWA/RD-86/168, Prefabricated Vertical Drains, (Rixner, et
al., 1986)

Blast Densification:
Blast Densification for Mitigation of Dynamic Settlement and Liquefaction,
Kimmerling, R. E., 1994, WSDOT Research Report WA-RD 348.1
Soil Improvement: State-of-the-Art Report, Mitchell, J. K., 1981, Proceedings
of the 10th International Conference on Soil Mechanics and Foundation
Engineering, Stockholm, Sweden, pp. 509-565.

Lime and Cement Soil Treatment:


Alaska DOT/FHWA Report No. FHWA-AK-RD-01-6B, Alaska Soil
Stabilization Design Guide, (Hicks, 2002)

Volume 1
3

ODOT Geotechnical Design Manual


April 2010

References

Barkdale, R. D., and Bachus, R. C., 1983, Design and Construction of Stone Columns Vol. 1,
Federal Highway Administration, FHWA/RD-83/02C.
Elias, V., Welsh, J., Warren, J., and Lukas, R., 2000, Ground Improvement Technical Summaries
Vol. 1 and 2, Demonstration Project 116, Federal Highway Administration, FHWA-SA-98-086.
Hicks, R. G., 2002, Alaska Soil Stabilization Design Guide, Alaska Department of Transportation and
Federal Highway Administration Report No. FHWA-AK-RD-01-6B.
Kimmerling, R. E., 1994, Blast Densification for Mitigation of Dynamic Settlement and Liquefaction,
WSDOT Research Report WA-RD 348.1, 114 pp.
Lukas, R. G., 1995, Geotechnical Engineering Circular No 1 - Dynamic Compaction, Federal
Highway Administration, FHWA-SA-95-037.
Mitchell, J. K., 1981, Soil Improvement: State-of-the-Art Report, Proceedings of the 10th International
Conference on Soil Mechanics and Foundation Engineering, Stockholm, Sweden, pp. 509-565.
Rixner, J. J., Kraemer, S. R., and Smith, A. D., 1986, Prefabricated Vertical Drains - Vol. :
Engineering Guidelines, Federal Highway Administration, FHWA/RD-86/168.

Volume 1
4

ODOT Geotechnical Design Manual


April 2010

Chapter

12

Rock Cuts Analysis, Design and


Mitigation

General

This chapter discusses the analysis, design guidelines and standards for rock slopes adjacent to
highways. Rock slope design for material sources is discussed in Chapter 20.

ODOT Rock Slope Design Policy

The purpose of the policy is to establish slope design standards for rock cuts and to encourage the
active involvement of geologists and geotechnical engineers in the rock slope design process. This
involvement is intended to ensure that rock slopes are safe to construct and economical and will
optimize safety for the public. In general, the policy includes four sections that deal with rock slopes.
These sections cover the rockslope design, rock fallout area requirements, the use of benches, and
rock slope stabilization and mitigation techniques.

12.2.1

Rock Slope Design

The purpose of the rock slope design is to develop rock cuts that will be safe to construct and will
provide long term safety for the public. The inclination of rock slopes should be based on the
structural geology and stability of the rock units, as described in the Geology or Geotechnical Report.
Rock unit slopes of vertical, 0.25:1, 0.5:1, 0.75:1 and 1:1 are commonly considered. The design rock
cut slope should be the steepest continuous slope (without benches) that satisfies physical and
stability considerations. Controlled blasting (using presplit and trim blasting techniques) is normally
required for rock cut slopes from vertical to 0.75:1. The purpose of controlled blasting is to minimize
blast damage to the rock backslope to help insure long-term-stability, improve safety, and lessen
maintenance. See Section 12.5 for more details regarding rock slope design.

12.2.2

Rockslope Fallout Areas

Fallout areas should be used where hazardous rockfall could occur. The fallout area is a nontraveled area between the highway and the cutslope with minimum width, depth and slope
Volume 1
1

ODOT Geotechnical Design Manual


April 2010

requirements. The minimum dimensions should be determined based on rock cut slope inclination
and height. The depth of the fallout area varies with the slope configuration. A preliminary
determination of the fallout area or catch ditch dimensions can be obtained from the Ritchie Rockfall
Catch Ditch Design Chart located in the ODOT Highway Design Manual, section 10.4.
Final catch ditch dimensions should be determined using the Rockfall Catchment Area Design Guide
(FHWA Final Report SPR-03(032).
As noted in the 2003 ODOT Highway Design Manual, section 10.4.4, a goal of 90% retention of rock
in the catchment area has been adopted for all new and reconstructed rock slopes. This goal may
not be achievable in all cases due to cost, environmental reasons, or other factors. The catchment
area depth may be achieved in a number of ways, including excavation and/or placing suitable
retaining structures at the highway shoulder. Where the slopes are inclined at flatter than 0.75:1, and
where the anticipated size of a single rock is less than 2 feet in diameter, chain link catch fences may
be considered as a substitute for depth of fallout. Slopes less than 40 feet high and flatter than 1:1
generally have a ditch and recoverable slope equal to or greater than a fallout ditch shown in the
Rockfall Catchment Area Design Guide. In that case, the standard roadway ditch will serve as
adequate rockfall catchment.
Temporary detours may require the construction of rockslopes and fallout areas. If the site has
previously been an area of rockfall activity, and the detour will reduce the fallout area, thereby putting
motorists in increased risk, the rockslope and fallout area must be designed to, at a minimum, not
increase the risk to the public. Fallout areas should then be designed to capture or retain at least as
much rockfall as was previously available prior to construction. Additional mitigation measures, along
with one way travel, reduced travel speed in the rockfall zone, and increased sight distances may be
required to reduce risk to the public. The designer should be prepared to address all of these issues
in the design process.

12.2.3

Benches

For most rock slope designs, benches should be avoided. The need for benches will be evaluated in
the geology and geotechnical investigations and described in the resulting reports. The minimum
bench design should satisfy the requirements outlined in the Rockfall Catchment Area Design Guide.
The bench configuration may be controlled by the need to perform periodic maintenance which
requires access to the bench. Soil and rock slopes may need a modification with benches to conform
to the environment or for safety and economic concerns. Following are some appropriate bench
applications.

Benching may improve slope stability where continuous slopes are not stable.

Where maintenance due to sloughing of soil overburden may be anticipated, a bench will
provide access and working room at the overburden rock contact.

Developing an access bench may facilitate construction where the top of cut begins at an
intermediate slope location.

On very high cuts, benches may be included for safety where rockfall is expected during
construction.

Where necessary, benches may be located to intercept and direct surface water runoff and
groundwater seepage to an appropriate collection facility.

All benches should be constructed to allow for maintenance access.

Volume 1
2

ODOT Geotechnical Design Manual


April 2010

12.2.4

Rock Slope Stabilization and Rockfall Mitigation


Techniques

Rock slope stabilization techniques may be required to accommodate special geologic features.
Stabilization techniques include rock bolts and dowels, wire mesh and cable net slope protection,
reinforced shotcrete, trim and production blasting. Specific stabilization techniques with appropriate
design will be recommended in the Geotechnical Report as necessary. Refer to Section 12.4 for
more detail.

Rockslope Stability Analysis

Slope stability analysis for rock slopes involves a thorough understanding of the structural geology
and rock mechanics. For most rock cuts on highway slopes, the stresses in the rock are much less
than the rock strength so there is little concern with the fracturing of the intact rock. Therefore,
stability is concerned with the stability of rock blocks formed by the discontinuities. Field data
collection of the dip, dip direction, nature and type of joint infilling, joint roughness and spacing are
important for the stability analysis of planar, wedge and toppling failure modes. Slope height, angle,
presence of potential rock launching features, block size, and block shape are important for the
analysis and design of rockfall mitigation techniques. Hand-calculation methods can be used to
analyze potential planar and wedge failures and computer programs such as ROCKPACK are also
available. Rockfall simulation programs, such as CRSP (Colorado Rockfall Simulation Program), are
used to analyze for rockfall catchment size and the prediction of rock kinetic energy. Only
geotechnical practitioners experienced in using these programs should perform the analysis. Refer
to Wyllie and Mah, 1998, for details on design, excavation, and stabilization of rock slopes.

Design Guidelines

General design guidelines are found in the references listed in Section 12.7. Design of rock slopes
adjacent to ODOT highways must also include consideration of additional factors such as
environmental issues, history of rockfall hazards, cost, risk/benefit, and needs of the project. The
following guidelines provide information on ODOT rock slope design.

12.4.1

Geologic Investigation and Mapping

For projects that include rock cuts, the geotechnical designer should contact the local Maintenance
district office to discuss the history of past rockfall events and consult the Region Geologist for the
project area to determine the RHRS (Rockfall Hazard Rating System) score and priority for that
highway and for the Region. The designer should also discuss the geologic hazard potential with the
Region Geologist so that a consensus on the degree of rockfall potential is reached. The discussions
will serve to highlight concerns regarding construction, local environmental needs, and feasible
options for mitigation of the hazard. The development and implementation of the geologic
investigation can then be completed.
Field data collection is generally done on a project site specific basis. Wiley and Mah, 1998,
discussed joint mapping techniques, stereographic projection, and types of subsurface exploration
that may be performed on rockslopes. Full scale tests of rockfall at the site may also be performed,
however, the cost and practicality of traffic control generally prevents this type of work.

12.4.2

Analysis and Design

As previously stated, analysis of planar, wedge and toppling failure modes can be performed by hand
or with some available computer programs. Wiley and Mah, 1998, discusses the analysis in detail.
Volume 1
3

ODOT Geotechnical Design Manual


April 2010

Simulation of rockfall using the CRSP computer program may be needed to determine the minimum
required dimensions of a rockfall catch ditch and the kinetic energy of rocks that may need to be
restrained by barriers, wire mesh, screens or walls. As a rule of thumb, draped gabion wire mesh
slope protection and screens are capable of withstanding impacts from rocks up to 2 feet in diameter.
For larger rocks, proprietary rockfall net systems or retaining walls will likely be needed. Experience
with the Rockfall Catchment Areas Design Guide study indicates that rockfall catch areas wider that
30 to 35 feet are not typically cost effective to construct, and additional barriers, fences or walls to
gain ditch depth become more cost effective than wider ditches.

12.4.3

Construction Issues

Construction of rock slopes near highways frequently must consider traffic control during blasting and
scaling operations. The traffic control may include adjacent railroad facilities where trains are running
next to the highway or other adjacent structures and facilities. The cost of traffic control for a busy
highway can potentially result in a doubling of the project cost. Therefore, careful consideration of
staging, detours, work zones and blast-produced flyrock control must be done during design. It may
even be necessary to choose another mitigation option than the preferred one because of these
issues.
Environmental concerns in scenic highway corridors have made construction of rock slopes more
difficult. Presplit hole half-casts that are visible after blasting may be regarded as a visual concern
and a bid item may be needed to partially or completely removed them. This issue has been most
notable in the Columbia River Gorge Scenic Corridor, and in a few USFS forest highways. Rock
coloration has also been a concern and a bid item for Permeon, a rock coloration product, has been
included on several projects contracts.

12.4.4

Blasting Consultant

A Blasting Consultant may need to be retained to assist a contractor in designing a safe blast if there
are nearby structures, if the site is particularly challenging, or otherwise has the potential to result in
undesired consequences. Guidelines for determining when a Blasting Consultant is needed are
located on the ODOT website. ODOT keeps a list of preapproved blasting consultants and has a
method of approving new blasting consultants and the HQ Geotechnical Group should be contacted.

12.4.5

Gabion Wire Mesh Slope Protection/ Cable Net Slope


Protection

For gabion wire mesh slope protection, the designer must choose either galvanized or PVC coated
wire in order to place the correct Standard Detail in the construction plans. Anchor spacing for Wire
Mesh, Cable Net, and Post-Supported Wire Mesh Slope Protection are based on the weight of the
mesh alone. Narrower spacing may be required where snow and ice loads will add a significant
amount of stress to the anchors.
The WashDOT research report, Design Guidelines for Wire Mesh/Cable Net Slope Protection, WARD 612.1, should be used to determine anchor spacing in snow/ice load situations. If mesh is use in
a coastal environment, stainless steel fasteners and hardware or heavy galvanizing should be used
to inhibit corrosion.

12.4.6

Rock Reinforcing Bolts and Rock Reinforcing Dowels

The designer must identify the installation area, size and strength of steel, pattern or spacing,
inclination, minimum length, and design loads of the bolts or dowels and this information must be
Volume 1
4

ODOT Geotechnical Design Manual


April 2010

included in the Geotechnical Report. Since rock reinforcing bolts are considered to be permanent,
acceptable materials for bonding the bolt into rock are non-shrink cement grouts, while polyester
resin or cement grout is acceptable for the semi-permanent rock reinforcing dowels. Mechanical
anchorage bolts and non-shrink cement grout are included in the ODOT Qualified Products List
(QPL). Split set and bail set type anchorage systems are considered temporary or low stress
installations and are not acceptable for use on ODOT projects.

12.4.7

Proprietary Rockfall Net Systems

High capacity rockfall net systems are available from two accepted manufacturers, GeoBrugg and
ROTEC International. Full scale tests on these systems have been performed by the manufacturers.
The systems are generally capable of withstanding impact kinetic energies up to 735 ft-tons and can
be constructed with breakaway post base connections and post heights up to 20 to 25 feet. These
systems are expensive and can raise objections about their highly visible nature. However, they can
be a viable alternative to high barriers and MSE walls in rockfall situations.

Standard Details

The ODOT GeoEnvironmental webpage includes a link to Standard Details normally used in the
mitigation of rockfall hazards. These details are also found in the Roadway Contract Plans
Development Guide. The following details, in English and Metric units, are presented:

Det 2200 - Cable Net Slope Protection Detail

Det 2201 - Wire Mesh/Cable Net Anchors Detail

Det 2202 - Shotcrete Slope Detail

Det 2203 - Wire Mesh Slope Protection Detail

Det 2204 - Barrier Mounted Rock Protection Screen Detail

Det 2205 - Post Supported Wire Mesh Slope Protection Detail

Det 2206 - Post Supported Wire Mesh Slope Protection Detail

Det 2207 - Post Supported Wire Mesh Slope Protection and Post Supported Rock
Protection Screen Anchor Details

Det 2208 - Rock Protection Screen Behind Concrete Barrier or Guardrail Detail

Det 2209 - Rock Protection Screen Behind Concrete Barrier or Guardrail Detail

Specifications

The location of Standard Specifications and Special Provisions for items pertaining to rockslopes and
rockslope mitigation are listed in the next sections.

Volume 1
5

ODOT Geotechnical Design Manual


April 2010

12.6.1

Blasting

Specifications for general excavation of rock slopes flatter than 0.75:1, where presplit (controlled
blasting) of the backslope is not required, are located in Section 00330.41(e) - Blasting of the
Standard Specifications.
Specifications for rock excavation where slopes are 0.75:1 or steeper are located in Section 00335 Blasting Methods and Protection of Excavation Backslopes of the Standard Specifications. A per foot
bid item quantity for Controlled Blast Holes is required if this specification is used.
Special Provisions for retaining a Blasting Consultant (see Section 00335.44 Blasting Consultant),
Vibration Control (see Section 00335.45 Vibration Control), and Blasting Noise Control (see Section
00335.46 Airblast and Noise Control) are located in the Special Provisions section of the ODOT
Specifications Webpage.

12.6.2

Rockslope Mitigation Methods

The following rockslope mitigation methods are located in a new section of the Standard
Specifications, Section 00398 Rockslope Stabilization and Reinforcement.

Wire Mesh Slope Protection

Post Supported Wire Mesh Slope Protection

Rock Protection Screen Behind Barrier or Guardrail

Barrier Mounted Rock Protection Screen

Rock Reinforcing Bolts/Rock Reinforcing Dowels

Proprietary Rockfall Net System

Volume 1
6

ODOT Geotechnical Design Manual


April 2010

References

Wyllie, D., and Mah, C., 1998, Rock Slopes Reference Manual, FHWA HI-99-007.
Pierson, L., Gullixson, C., and Chassie, R., 2001, Rockfall Catchment Area Design Guide, FHWA
Final Report SPR-3(032).
Konya, C., and Walter, E., 2003, Rock Blasting and Overbreak Control, 2nd ed., FHWA-HI-92-001.
Muhunthan, B. et.al, 2005, Analysis and Design of Wire Mesh/Cable Net Slope Protection,
WashDOT, Final Research Report WA-RD 612.1.
Turner, A. K., and Schuster, R.L., 1996, Landslides: Investigation and Mitigation, TRB Special Report
247, National Academy Press.
Brawner, C. O., 1993, Manual of Practice on Rockfall Hazard Mitigation Methods, FHWA, National
Highway Institute Participant Workbook DTFH61-92-Z-00069.
Post Tensioning Institute, 1996, Recommendations for Prestressed Rock and Soil Anchors, Post
Tensioning Institute, Phoenix, Arizona.
Jones, C., Higgins, J., and Andrew, R., 2000, CRSP ver. 4.0, Colorado Rockfall Simulation
Program, Colorado DOT Report CDOT-SYMB-CGS-99-1.
Watts, C. F., 1996, ROCKPACK II, Rock Slope Stability Computerized Analysis Package, Users
Manual, C.F. Watts Assoc., Radford VA.

Volume 1
1

ODOT Geotechnical Design Manual


April 2010

Chapter

13

Slope Stability Analysis

General

This Chapter to be completed at a later date.

Volume 1
1

ODOT Geotechnical Design Manual


April 2010

Chapter

14

Geosynthetic Design

General

This Chapter to be completed at a later date.

Volume 1
1

ODOT Geotechnical Design Manual


April 2010

You might also like